content
stringlengths
1
15.9M
\section{Introduction} Doping of antiferromagnetic Mott insulators causes frustration which has a profound effect on the magnetic properties of such systems. High temperature superconductivity, charge ordering and anomalous transport properties have been observed in doped transition metal oxides with perovskite structure (see Ref.~\cite{dagx:99} for a review). Numerical studies of one-dimensional models proposed for doped Nickel-oxides and Manganites for such systems showed strong tendencies toward ferromagnetism and phase separation. For a better understanding of these phenomena requires to take into account strong electronic correlations. A commonly used starting point for a description of such systems is the ferromagnetic Kondo Hamiltonian \cite{Zener51,AnHa55} \begin{equation} {\cal H} = -t \sum_{\langle ij\rangle\sigma} \left(c^\dagger_{i\sigma} c_{j\sigma} + h.c.\right) - J_H \sum_i {{\sigma}}_i\cdot \mathbf{S}_i \label{Hel} \end{equation} where the first term is the kinetic energy given in terms of canonical fermionic creation and annihilation operators $c^\dagger_{i\sigma}$ and $c_{i\sigma}$ for the itinerant electrons on site $i$ with spin $\sigma=\uparrow,\downarrow$ and the second is the ferromagnetic Hund's rule coupling between the spins ${\sigma}_i^k = \sum_{\alpha\beta} c^\dagger_{i\alpha} \left({\sigma}^k \right)_{\alpha\beta} c_{i\beta}$, $k=x,y,z$, of the itinerant electrons to a localized spin $\mathbf{S}_n$. A Coulombic repulsion to suppress double occupancy in the itinerant band is implicit. Large Hund coupling $J_H$ favours the alignment of the itinerant and the localized spin, i.e.\ spin eigenstates with the maximum allowed total spin. Hence, for local spins of length $S-1/2$ the degrees of freedom that need to be kept for an effective theory of the low lying modes in this model are spin-$S$ ``spins'' and spin-$S-1/2$ ``holes''. Within the double-exchange Hamiltonian \cite{Zener51,AnHa55} classical ``background spins'' $\mathbf{S}_i$ are used to approximate these holes. However, the non trivial phases arising in a full quantum mechanical treatment of these spin degrees of freedom may induce low-energy modes which are essential for an understanding of the magnetic properties of these systems \cite{mhda:96}. To obtain an effective lattice model on the $4S+1$ dimensional local Hilbertspace one has to eliminate the other allowed spin configurations in a perturbative analysis for $J_H\gg t$ \cite{dagx:96,mhda:96}. To leading order in $J_H$ the effective Hamiltonian of the resulting model is simply the projection of (\ref{Hel}) onto the states listed above. $SU(2)$-invariance implies that this operator can be written as a polynomial of spin-operators: \begin{equation} {\cal H}_{\rm eff} \approx -t \sum_{\langle ij\rangle} {\cal P}_{ij} Q_S(y_{ij})\, \quad y_{ij} = \mathbf{S}_i\cdot\mathbf{S}_j/S(S-1/2) \label{Heleff} \end{equation} where the operator ${\cal P}_{ij}$ permutes the states on sites $i$ and $j$ thereby allowing the holes to move. For large but finite $J_H$ one finds additional antiferromagnetic Heisenberg exchange terms. Hamiltonian operators of this type have been used as a starting point for studies the phase diagram of doped transition metal oxides by numerical diagonalization of small clusters \cite{dagx:96,riera:97,dagx:99}. In this paper we introduce a class of exactly solvable models in one spatial dimension which generalize the supersymmetric $t$--$J$ model \cite{lai:74,suth:75,schl:87} and a model for doped spin-$1$ chains \cite{fpt:98}. In spite of the appearance of several additional couplings guaranteeing integrability these models may provide further insights into the peculiar properties of these compounds. In the following section we shall use the framework of the Quantum Inverse Scattering Method (QISM) to construct families of vertex models making use of so called `atypical' representations of the super Lie algebra $gl(2|1)$. The spectra of the corresponding commuting transfer matrices are obtained by means of the algebraic Bethe Ansatz (BA). In Sect.~\ref{sec:fusion} we derive local Hamiltonians (i.e.\ operators involving nearest neighbour interactions on the lattice only) similar to the ones discussed above using a fusion procedure. In Sec.~\ref{sec:tba} integral equations determinining the spectra and thermodynamic properties of these models are obtained from the BA equations. From these equations we obtain the phase diagram of the doped spin chains in a magnetic field for low temperatures $T\ll H$ and the low temperature properties for vanishing field $H=0$ in Sects.~\ref{sec:phasH} and \ref{sec:phasT}. We conclude with some remarks on a possible $SU(2)$-invariant effective field theory description for the low energy/low temperature sector of these systems. \section{Construction of the models} \label{sec:constr} Below we will construct a class of integrable vertex models from solutions to the Yang Baxter equation (YBE) which are invariant under the action of the graded Lie algebra $gl(2|1)$ \cite{snr:77,marcu:80}. The nine generators of $gl(2|1)$ are classified into even ($1$, $S^z$, $S^\pm$, $B$) and odd ($V_\pm$, $W_\pm$) ones depending on their parity w.r.t.\ grading. The even generators are the spin operators $S^\alpha$ form a $SU(2)$ subalgebra with commutation relations $[S^z,S^{\pm}]\,=\,\pm S^{\pm}$, $[S^+,S^-]\,=\,2S^z$ and the $U(1)$ charge $B$ which commutes with the $S^\alpha$: $[B,S^{\pm}]\,=\,[B,S^z]\,=\,0$. The commutators between even and odd generators of the algebra are \begin{eqnarray} \left[S^z,V_{\pm}\right]\,=\,\pm\frac{1}{2}V_{\pm},\quad \left[S^{\pm},V_{\pm}\right]\,=\,0, && \left[S^{\mp},V_{\pm}\right]\,=\,V_{\mp},\quad \left[B,V_{\pm}\right]\,=\,\frac{1}{2}V_{\pm}, \nonumber\\ \left[S^z,W_{\pm}\right]\,=\,\pm\frac{1}{2}W_{\pm},\quad \left[S^{\pm},W_{\pm}\right]\,=\,0, && \left[S^{\mp},W_{\pm}\right]\,=\,W_{\mp},\quad \left[B,W_{\pm}\right]\,=\,-\frac{1}{2}W_{\pm}, \label{comm:eo} \end{eqnarray} and the odd generators satisfy \emph{anti}commutation rules \begin{eqnarray} &&\{V_\pm,V_\pm\} = \{V_\pm,V_\mp\} = \{W_\pm,W_\pm\} = \{W_\pm,W_\mp\}=0 \ , \nonumber\\ &&\{V_\pm,W_\pm\} = \pm \frac{1}{2} S^\pm\ , \quad \{V_\pm,W_\mp\} = \frac{1}{2}\left(-S^z \pm B\right) . \label{comm:oo} \end{eqnarray} The 'typical' representations $[b,s]$ of this algebra can be characterized by the eigenvalues of operators $B$ and $\mathbf{S}^2$ on the multiplet with largest total $SU(2)$-spin \cite{snr:77,marcu:80}. Their dimension is $8S$ and they can be decomposed into two spin-$(S-{1/2})$ multiplets with charge $b\pm1/2$ and a spin-$S$ and a spin-$(S-1)$ multiplet with charge $b$ with respect to the $SU(2)$-subalgebra of $gl(2|1)$. In the following we shall be particularly interested in the $(4S+1)$-dimensional so-called 'atypical' representations $[S]_+$ which contain two multiplets of spin $S$ and $(S-{1/2})$ and corresponding charges $b=S$ and $b=S+1/2$. Choosing a basis $\{|b,s,m\rangle\}$ in which $B$, $\mathbf{S}^2$ and $S^z$ are diagonal, the nonvanishing matrix elements of the fermionic operators are \begin{eqnarray} \langle S+{1\over2},S-{1\over2},m\pm{1\over2}| V_\pm |S,S,m\rangle &=& \pm \sqrt{\frac{S\mp m}{2}} \nonumber\\ \langle S,S,m| W_\pm |S+{1\over2},S-{1\over2},m\mp{1\over2}\rangle &=& \sqrt{\frac{S\pm m}{2}}\ . \label{repVW} \end{eqnarray} Tensor products of atypical representation can be decomposed as \begin{eqnarray} && \left[S\right]_+ \otimes \left[S'\right]_+ = \left[S+S'\right]_+ \oplus \left[S+S'+{1\over2},S+S'-{1\over2}\right] \nonumber\\ &&\qquad \oplus \left[S+S'+{1\over2},S+S'-{3\over2}\right] \oplus\cdots \oplus \left[S+S'+{1\over2},|S-S'|+{1\over2}\right]\, . \label{tensor} \end{eqnarray} The irreducible components in this tensor product can be identified by the action of the quadratic Casimir of the algebra \begin{equation} K_2 = {\mathbf S}^2 - B^2 - W_-V_+ + W_+V_- - V_-W_+ + V_+W_-\, \label{casimir2} \end{equation} which has eigenvalues 0 on $\left[s\right]_+$ and $(s^2-b^2)$ on $\left[b,s\right]$. Choosing an irreducible $d$-dimensional representation of $gl(2|1)$ acting on a quantum space ${\cal V}\sim \mathbb{C}^d$, it is straightforward to verify that the ${\cal L}$-operator \cite{kulish:85} \begin{equation} {\cal L}(\mu) = \left( \begin{array}{ccc} \mu+ 2iB & i\sqrt{2}W_- & i\sqrt{2} W_+ \\ i\sqrt{2}V_+ & \mu+i(B+S^z) & -iS^+ \\ -i\sqrt{2}V_- & -iS^- & \mu+i(B-S^z) \end{array}\right)\ . \label{lop} \end{equation} written as a matrix in the three-dimensional matrix space ${\cal M}$ solves the Yang-Baxter equation \begin{equation} R(\lambda-\mu) \left({\cal L}(\lambda)\otimes{\cal L}(\mu)\right) =\left({\cal L}(\mu)\otimes{\cal L}(\lambda)\right) R(\lambda-\mu) \label{ybe3S} \end{equation} with the $R$-matrix \begin{equation} R(\lambda)=b(\lambda)I + a(\lambda) \Pi\ ,\quad a(\lambda)=\frac{\lambda}{\lambda+i}\ ,\quad b(\lambda)=\frac{i}{\lambda+i}\ . \label{rmat33} \end{equation} Here $I$ and $\Pi$ are the unit and \emph{graded} permutation operator acting on the tensor product ${\cal M}_1\otimes{\cal M}_2$ of matrix spaces in which ${\cal L}$-operators act in (\ref{ybe3S}). Assigning Grassmann parities $\epsilon_i\in\{0,1\}$ to the basis of these spaces the matrix elements of $\Pi$ are \begin{equation} \Pi_{i_2,j_2}^{i_1,j_1} = (-1)^{\epsilon_{j_1}\epsilon_{j_2}} \delta_{i_1j_2}\delta_{i_2j_1}\ . \end{equation} Similarly, the matrix elements of the operators acting on the tensor product of these spaces pick up signs \( \left(A\otimes B\right)_{i_2j_2}^{i_1j_1} = (-1)^{\epsilon_{i_2}(\epsilon_{i_1}+\epsilon_{j_1})} A_{i_1j_1} B_{i_2j_2} \) due to the grading of the basis. Considering the ${\cal L}$-operator as a linear operator acting on the tensor-product of spaces ${\cal M}\otimes {\cal V}$ with the fundamental three-dimensional representation $\left[1/2\right]_+$ acting on the matrix space its $gl(2|1)$-invariance can be established by rewriting (\ref{lop}) as $\mu-iK_2$ in terms of the Casimir operator (\ref{casimir2}) on the tensor product (up to a shift of the spectral parameter). The intertwining relation (\ref{ybe3S}) implies that the monodromy matrix, defined as the matrix product \begin{equation} {\cal T(\lambda)} = {\cal L}_L(\lambda){\cal L}_{L-1}(\lambda) \cdots{\cal L}_1(\lambda) \label{mono33} \end{equation} of ${\cal L}_n$-operators (\ref{lop}) with entries acting on different quantum spaces ${\cal V}_n$ satisfies a Yang-Baxter equation with the same $R$-matrix (\ref{rmat33}): \begin{equation} R(\lambda-\mu) \left({\cal T}(\lambda)\otimes{\cal T}(\mu)\right) =\left({\cal T}(\mu)\otimes{\cal T}(\lambda)\right) R(\lambda-\mu)\ . \label{ybeT3S} \end{equation} As an immediate consequence of this identity the transfer matrix given by the matrix super trace of ${\cal T}$ \begin{equation} t_{3}(\mu) = sTr\left({\cal T}(\mu)\right) =\sum_{i=1}^3 (-1)^{\epsilon_i} \left[ {\cal T}(\mu) \right]^{ii} \label{trans3S} \end{equation} commutes for different values of the spectral parameter $\mu$ thus being the generating functional for a family of commuting operators on the graded tensor product of $L$ quantum spaces which we will identify below with the Hilbert space of an integrable spin chain. The subscript to the transfer matrix is used to label the dimension of the matrix space of the corresponding monodromy matrix. The spectrum of this transfer matrix is obtained by means of the algebraic Bethe Ansatz (ABA) \cite{vladb}. As a consequence of the grading different sets of Bethe Ansatz equations (BAE) follow from different orderings of the basis \cite{kulish:85}. We now restrict ourselves to representations $\left[S\right]_+$ in the quantum spaces ${\cal V}_n$ where we choose the state $|0\rangle_n\equiv|S,S,S\rangle_n$ as our reference state. The two other possible sets of BAE for this model are given in Appendix~\ref{app:ba}, their equivalence is shown in Appendix~\ref{app:eq}. Since the ABA for the transfer matrix (\ref{trans3S}) is completely analogeous to the case considered in \cite{kulish:85,esko:92,foka:93}, we only sketch the main steps leading to the BAE and the spectrum. The action of (\ref{lop}) on the reference state is \begin{equation} {\cal L}_n(\mu)|0\rangle_n = \left(\begin{array}{ccc} \mu+2iS & 0 & 0 \\ 0 & \mu+2iS & 0 \\ -i\sqrt{2}V_n^- & -iS_n^- & \mu \end{array}\right)|0\rangle_n\ . \end{equation} Similarly, acting with the monodrony matrix (\ref{mono33}) on the state $|\Omega_S\rangle = |0\rangle_L\otimes\cdots\otimes|0\rangle_1$ we get \begin{equation} {\cal T}(\mu)|\Omega_S\rangle = \left(\begin{array}{ccc} (\mu+2iS)^L & 0 & 0 \\ 0 & (\mu+2iS)^L & 0 \\ C_1(\mu) & C_2(\mu) & \mu^L \end{array}\right)|\Omega_S\rangle\ . \end{equation} Hence, $|\Omega_S\rangle$ is an eigenstate of the transfer matrix (\ref{trans3S}) with eigenvalue $(-\mu^L)$ (we have chosen the grading $\epsilon_1=0$, $\epsilon_2=\epsilon_3=1$ in the matrix space here). The operators $C_1(\lambda)$ and $C_2(\lambda)$ create a hole and lower the spin of the system respectively. For eigenstates of $t_{3S}(\lambda)$ with $N_h$ holes (generating sites with spin $S-{1/2}$) and magnetization $M^z=LS -{1\over2}N_h - N_\downarrow$ we make the Ansatz \begin{equation} |\tilde\lambda_1,\ldots,\tilde\lambda_n|F\rangle = C_{a_1}(\tilde\lambda_1)\cdots C_{a_n}(\tilde\lambda_n) |\Omega_S\rangle F^{a_n\cdots a_1} \label{BAvec} \end{equation} where $n=N_h+N_\downarrow$. Using the algebra of the operators in (\ref{ybeT3S}) we are led to an eigenvalue problem for the amplitudes $F^{a_n\cdots a_1}$ which is solved by a second Bethe Ansatz parametrized by 'hole rapidities' $\{\tilde\nu_\alpha\}_{\alpha=1}^{N_h}$. Finally, we find that (\ref{BAvec}) is an eigenstate of (\ref{trans3S}) with eigenvalue \begin{eqnarray} &&\Lambda_{3}\left(\mu| \{\tilde\lambda_j\}_{j=1}^{N_h+N_\downarrow}, \{\tilde\nu_\alpha\}_{\alpha=1}^{N_h}\right) = -{\mu}^L \prod_{j=1}^{N_h+N_\downarrow} {\mu-\tilde\lambda_j+i \over \mu-\tilde\lambda_j} \nonumber\\ &&\qquad +\left({\mu+2iS}\right)^L\prod_{\alpha=1}^{N_h} {\mu-\tilde\nu_\alpha +i \over \mu-\tilde\nu_\alpha} \left\{ 1 - \prod_{j=1}^{N_h+N_\downarrow} {\tilde\lambda_j-\mu+i \over \tilde\lambda_j-\mu} \right\} \end{eqnarray} provided the spectral parameters $\tilde\lambda_j\equiv\lambda_j -iS$ and $\tilde\nu_\alpha\equiv\nu_\alpha-iS+i/2$ satisfy the following set of BAE \begin{eqnarray} \left( {\lambda_j+iS\over \lambda_j-iS} \right)^L &=& \prod_{k\ne j}^{N_h+N_\downarrow} {{\lambda_j-\lambda_k+i}\over{\lambda_j-\lambda_k-i}}\, \prod_{\alpha=1}^{N_h} {{\lambda_j-\nu_\alpha-{i\over2}} \over {\lambda_j-\nu_\alpha+{i\over2}} }\ , \nonumber\\ && \qquad j=1,\ldots,N_h+N_\downarrow \label{bae3}\\ 1 &=& \prod_{k=1}^{N_h+N_\downarrow} {{\nu_\alpha - \lambda_k +{i\over2}} \over {\nu_\alpha - \lambda_k -{i\over2}}}\ , \quad \alpha=1,\ldots,N_h\ . \nonumber \end{eqnarray} \section{Doped spin chains} \label{sec:fusion} Choosing the fundamental three dimensional representation $\left[1/2\right]_+$ of $gl(2|1)$ in (\ref{lop}), the ${\cal L}$-operator taken at $\mu=-i$ becomes a graded permutation operator on the tensor product ${\cal M}\otimes{\cal V}$ of matrix and quantum space. Hence the transfer matrix (\ref{trans3S}) generates a translation by one site on the lattice for this value of the spectral parameter. Expanding the logarithm of $t_3(\mu)$ about this shift point we can construct a Hamiltonian with nearest neighbour interactions only (\ref{trans3S}) \begin{equation} -i\left.{\partial\over\partial\mu} \ln t_{3}(\mu)\right|_{\mu=-i} = \sum_{n=1}^L \ldots \label{Hstj} \end{equation} which is the supersymmetric $t$--$J$ model (see e.g.\ \cite{esko:92}). In this case, Eqs.~(\ref{bae3}) are known as Sutherland's form of the BAE for this model \cite{suth:75}. For representation different from $\left[{1/2}\right]_+$ it is not possible to construct a local Hamiltonian directly from the ${\cal L}$-operators (\ref{lop}) (they can be used to construct $t$--$J$ models perturbed by integrable impurities though \cite{bef:96,sczv:97,schl:98}). To obtain a homogenous lattice model such as (\ref{Hstj}) new ${\cal L}$-operators have to be found which act on tensor products of matrix and quantum spaces of the same dimension. Their $gl(2|1)$-invariance implies that they can be expressed as sum over multiples of the projectors on irreducible components of the tensor product of representations in the two spaces. Noting that the \emph{spin} multiplets at charge $(S+S')$ in the tensor product (\ref{tensor}) are just the ones obtained by adding two spins of length $S$, $S'$ (and similarly spins $(S-1/2)$, $(S'-1/2)$ at charge $(S+S'+1)$) we find linear operators ${\cal L}^{\{SS'\}} (\mu)$ acting on spaces carrying atypical representations $\left[S\right]_+$ and $\left[S'\right]_+$ from the $gl(2)$-invariant ones constructed in Ref.~\cite{kulx:81}, namely: \begin{equation} {\cal L}^{\{SS'\}}(\mu) = - \prod_{k=|S-S'|+1}^{S+S'} {{\mu-i k}\over{\mu+ik}}\, {\cal P}_{\left[S+S'\right]_+} - \sum_{m=|S-S'|}^{S+S'-1} \,\, \prod_{k=|S-S'|+1}^m {{\mu-i k}\over{\mu+ik}}\, {\cal P}_{\left[S+S'+{1\over2},m+{1\over2}\right]}\, . \label{lopss} \end{equation} Here ${\cal P}_\Lambda$ is a projector on the $gl(2|1)$-multiplet $\Lambda$ in the tensor product $\left[S\right]_+\otimes \left[S'\right]_+$. Choosing one of the representations to be $\left[1/2\right]_+$ and comparing this expression with (\ref{lop}) we find \begin{equation} {\cal L}^{\{1/2,S\}}\left(\mu\right) = -{1\over\mu+i(S+1/2)}\, {\cal L}\left(\mu-i(S+1/2)\right)\, . \end{equation} The new ${\cal L}$-operators satisfy the YBEs \begin{equation} R_{S_1S_2}(\lambda-\mu) \left({\cal L}^{\{S_1S_3\}}(\lambda) \otimes{\cal L}^{\{S_2S_3\}}(\mu)\right) = \left({\cal L}^{\{S_2S_3\}}(\mu) \otimes{\cal L}^{\{S_1S_3\}}(\lambda)\right) R_{S_1S_2}(\lambda-\mu) \label{ybeSSS} \end{equation} where $R_{SS'}=\Pi{\cal L}^{\{SS'\}}$. As a consequence of this relation the transfer matrices of the corresponding vertex models \begin{equation} t^{\{S_0S\}}(\mu) = sTr_0\left({\cal L}_L^{\{S_0S\}}(\mu) \cdots {\cal L}_1^{\{S_0S\}}(\mu)\right) \label{transSSS} \end{equation} (the product of ${\cal L}$-operators and the super trace are taken in the $(4S_0+1)$-dimensional matrix space) commute, i.e. $\left[ t^{\{S_0S\}}(\mu), t^{\{S_1S\}}(\lambda) \right] =0$ for all $S_0$, $S_1$. From (\ref{lopss}) we observe that ${\cal L}^{SS}(\mu=0) \propto \Pi$. As in (\ref{Hstj}) an integrable Hamiltonian with nearest neighbour interactions on the lattice can be constructed by taking the logarithmic derivative of $\ln t^{\{SS\}}(\mu)$ at $\mu=0$. The eigenstates of the transfer matrices $t^{\{S'S\}}(\mu)$ are parametrized by the roots of the BAE (\ref{bae3}). To compute their eigenvalues we need the so-called fusion relations between these operators for different $S'$ which are obtained from considering tensor products of different matrix spaces. Starting with the YBE (\ref{ybeSSS}) for $S_1=S_2=1/2$ we observe that choosing $\lambda-\mu=-i$ the matrix $R_{{1\over2},{1\over2}}$ is proportional to a projector onto the five-dimensional subrepresentation $\left[1\right]_+$ of the tensor product $\left[1/2\right]_+ \otimes \left[1/2\right]_+$. This implies that the ${\cal L}$-operator $\tilde{\cal L}(\mu)= {\cal L}^{\{1/2,S\}} \left(\mu-i/2\right) \otimes {\cal L}^{\{1/2,S\}} \left(\mu+i/2\right)$ satisfies the condition \begin{equation} {\cal P}_{\left[1\right]_+} \tilde{\cal L}(\mu) {\cal P}_{\left[3/2,1/2\right]} \equiv 0\ . \end{equation} Consequently, it can be rewritten as \begin{equation} \tilde{\cal L}(\mu) \sim \left(\begin{array}{cc} {\cal L}^{\{\left[3/2,1/2\right],S\}}(\mu) & * \\ 0 & {\cal L}^{\{1,S\}}(\mu) \end{array}\right) \label{lopfus} \end{equation} after a proper reordering of the basis in ${\cal M}_1\otimes{\cal M}_2$. Here ${\cal L}^{\{\left[3/2,1/2\right],S\}}(\mu)$ is a $4\times4$ matrix acting on the $\left[3/2,1/2\right]$ component of this tensor product. Building a monodromy matrix from $L$ copies of (\ref{lopfus}) we obtain the fusion relation for the corresponding transfer matrices \begin{eqnarray} \tilde{t}(\mu) &\equiv& sTr\left( \tilde{\cal L}_L(\mu) \ldots \tilde{\cal L}_1(\mu) \right) \nonumber\\ &=& t^{\{1/2,S\}}\left(\mu-{i/2}\right) t^{\{1/2,S\}}\left(\mu+{i/2}\right) = t^{\{1,S\}}(\mu) + t^{\{\left[3/2,1/2\right],S\}}(\mu)\, \label{fuseq} \end{eqnarray} and an equivalent equation for their eigenvalues $\Lambda^{\{\cdot\,S\}} (\mu)$. Since there are still two unknown functions of $\mu$ on the RHS of this equation it is not possible to determine the spectrum of the new transfer matrices from (\ref{fuseq}) alone. As additional information we make use of the fact that the eigenvalues of the transfer matrix are analytical functions of $\mu$, i.e.\ the residues at their simple poles vanish as a consequence of the BAE (\ref{bae3}). Complemented by the trivial action of $t^{\{1,S\}}(\mu)$ and $t^{\{\left[3/2, 1/2\right], S\}}(\mu)$ on the pseudo vacuum $|\Omega_S\rangle$ this allows to compute $\Lambda^{\{1,S\}}$ with the result (see e.g.\ \cite{pffr:96}) \begin{eqnarray} &&\Lambda^{\{1,S\}}\left(\mu| \{\lambda_j\}_{j=1}^{N_h+N_\downarrow}, \{\nu_\alpha\}_{\alpha=1}^{N_h}\right) = \left( {\mu-iS\over\mu+iS}\, {\mu-i(S+1)\over\mu+i(S+1)} \right)^L \prod_{j=1}^{N_h+N_\downarrow} {\mu-\lambda_j+i\over\mu-\lambda_j-i} \nonumber\\ &&\quad -\left( {\mu-iS\over\mu+iS}\, {\mu+i(S-1)\over\mu+i(S+1)} \right)^L \prod_{j=1}^{N_h+N_\downarrow} {\mu-\lambda_j+i\over\mu-\lambda_j} \prod_{\alpha=1}^{N_h} {\mu-\nu_\alpha-{i\over2}\over\mu-\nu_\alpha-{3i\over2}} \left\{1-\prod_{j=1}^{N_h+N_\downarrow} {\lambda_j-\mu+2i \over \lambda_j-\mu+i} \right\} \nonumber\\ &&\quad -\left( {\mu+i(S-1)\over\mu+i(S+1)} \right)^L \prod_{j=1}^{N_h+N_\downarrow} {\lambda_j-\mu+i\over\lambda_j-\mu} \prod_{\alpha=1}^{N_h} {\mu-\nu_\alpha+{i\over2} \over \mu-\nu_\alpha-{3i\over2}} \left\{1-\prod_{j=1}^{N_h+N_\downarrow} {\lambda_j-\mu+2i\over\lambda_j-\mu+i} \right\}\ . \label{Lambda5} \end{eqnarray} As observed above, the local vertex operator ${\cal L}^{\{1,1\}} (\mu=0)$ becomes a graded permutation operator on the tensor product of the five-dimensional matrix space and the quantum space in which the representation $\left[S=1\right]_+$ is acting. Hence we can proceed as for the the case of the fundamental representation above and obtain the local lattice Hamiltonian for a spin-1 chain doped with $S=1/2$ holes introduced in Ref.~\cite{fpt:98}: \begin{equation} {\cal H}^{(1)} = -i \left.{\partial\over\partial\mu} \ln\left(t^{\{1,1\}}(\mu)\right)\right|_{\mu=0} -3L =\sum_{n=1}^L \left\{ {\cal H}^{\rm exch}_{n,n+1} + {\cal H}^{\rm hopp}_{n,n+1}\right\}\, -N_h\, . \label{hamil1} \end{equation} The exchange and kinetic part of the Hamiltonian expressed in terms of spin operators ${\mathbf S_i}$ with ${\mathbf S}_i^2=S_i(S_i+1)$ with $S_i=1$ or ${1/2}$ read \begin{eqnarray} {\cal H}^{\rm exch}_{ij} &=& {1\over2}\left( {1\over S_i S_j} {\mathbf S}_i \cdot {\mathbf S}_j - 1 +\delta_{S_iS_j,1}\left(1-({\mathbf S}_i\cdot{\mathbf S}_j)^2\right) \right)\ , \nonumber\\ {\cal H}^{\rm hopp}_{ij} &=& -\left(1-\delta_{S_i,S_j}\right) {\cal P}_{ij} \left( {\mathbf S}_i \cdot {\mathbf S}_j \right)\ . \nonumber \end{eqnarray} ${\cal P}_{ij}$ permutes the spins on sites $i$ and $j$. The corresponding eigenvalues of (\ref{hamil1}) are obtained from (\ref{Lambda5}): adding an external magnetic field $H$ and a chemical potential \begin{eqnarray} && E^{(1)}\left(\{\lambda_j\},\{\nu_\alpha\}\right) -HM^z-\mu N_h \nonumber\\ && = \sum_{k=1}^{N_h+N_\downarrow} \left(H-{2\over\lambda_k^2+1}\right) -\sum_{\alpha=1}^{N_h} \left(\mu+{1\over2}H\right) -LH \label{heig} \end{eqnarray} (we have added an external magnetic $H$ field and a (hole) chemical potential $\mu$ to the Hamiltonian). To proceed to higher $S$ we iterate the procedure used above: for $S_1=1/2$ and $S_2=S'-1/2$ arbitrary we use the fact that $R_{{1\over2}, S'-{1\over2}} (\mu=-iS')\propto{\cal P}_{[S']_+}$ in the YBE (\ref{ybeSSS}). This leads to the fusion relation \begin{equation} t^{\{1/2,S\}}\left(\mu-i\left(S'-{1\over2}\right)\right)\, t^{\{S'-1/2,S\}}\left(\mu+{i\over2}\right) = t^{\{S',S\}}(\mu) + t^{\{\left[S'+1/2,S'-1/2\right],S\}}(\mu)\, \end{equation} which allows to determine the eigenvalues of $t^{\{S',S\}}(\mu)$ from the known ones of $t^{\{1/2,S\}}(\mu)$ and $t^{\{S'-1/2,S\}}(\mu)$ as \begin{eqnarray} \Lambda^{\{S',S\}}\left(\mu| \{\lambda_j\}_{j=1}^{N_h+N_\downarrow}, \{\nu_\alpha\}_{\alpha=1}^{N_h}\right) &=& \left( \prod_{k=|S-S'|+1}^{S+S'} {\mu-ik\over\mu+ik} \right)^L \prod_{j=1}^{N_h+N_\downarrow} {\mu-\lambda_j+iS'\over\mu-\lambda_j-iS'} \nonumber\\ && + \Big(\mu-i(S-S')\Big)^L \left\{\ldots\right\}\, . \end{eqnarray} The terms in braces are determined by the fusion equations together with the vanishing of the residues at the simple poles of $\Lambda^{\{S',S\}} (\mu)$ due to the BAE (\ref{bae3}). They do not contribute to the spectrum of the nearest neighbour spin chain Hamiltonian \begin{equation} {\cal H}^{(S)} = -i \left.{\partial\over\partial\mu} \ln\left(t^{\{S,S\}}(\mu)\right)\right|_{\mu=0} + {\rm const.} \label{HS} \end{equation} whose eigenvalues are \begin{eqnarray} && E^{(S)}\left(\{\lambda_j\}, \{\nu_\alpha\}\right) -HM^z-\mu N_h \nonumber\\ && = \sum_{k=1}^{N_h+N_\downarrow} \left(H-{2S\over\lambda_k^2+S^2}\right) -\sum_{\alpha=1}^{N_h} \left(\mu+{1\over2}H\right) -LH\ . \label{heigS} \end{eqnarray} \section{Thermodynamic Bethe Ansatz} \label{sec:tba} To study the thermodynamics of the doped spin chains (\ref{HS}) we have to analyze the BAE (\ref{bae3}). In the thermodynamic limit $L\to\infty$ general solutions are known to consist of real hole rapidities $\nu_\alpha$ and complex $n$-\emph{strings} of spin-rapidities \begin{equation} \lambda_j^{n,k} = \lambda_j^{(n)} + {i\over2}\left(n+1-2k\right)\, , \quad k=1,\ldots,n\, . \label{strings} \end{equation} Now we consider solutions of (\ref{bae3}) built from $N_h$ hole rapidities and $M_n$ $\lambda$-strings of length $n$. Rewriting the BAE in terms of the real variables $\nu_\alpha$ and $\lambda_j^n$ and taking the logarithm we obtain \begin{equation} z_c\left(\nu_\alpha\right) = {I_\alpha\over L}\ ,\qquad z_s^{(n)}\left(\lambda_j^{(n)}\right) = {J_j^{(n)}\over L}\ , \label{zaehl} \end{equation} where $J_j^{(n)}$ and $I_\alpha$ are integers (or half-odd integers) and the functions $z_i$ are given as \begin{eqnarray} {2\pi}z_s^{(n)}(\lambda) &=& \theta_{n,2S}\left({\lambda}\right) -{1\over L}\sum_{m=1}^\infty \sum_{j=1}^{M_{m}} \Xi_{nm}\left(\lambda-\lambda_j^{(m)}\right) + {1\over L}\sum_{\alpha=1}^{N_h} \theta_n\left({\lambda-\nu_\alpha}\right) \nonumber\\ {2\pi}z_c(\nu) &=& {1\over L} \sum_{n=1}^\infty \sum_{j=1}^{M_n} \theta_n\left({\nu-\lambda_j^{(n)}}\right) \label{zz} \end{eqnarray} where $\theta_n(x) = 2\arctan(2x/n)$ and \begin{eqnarray} \theta_{nm}(x) &=& \theta_{m+n-1}\left(x\right) + \theta_{m+n-3}\left(x\right)+\ldots + \theta_{|m-n|+1}\left(x\right)\ , \nonumber\\ \Xi_{nm}(x) &=& \theta_{n+m}\left({x}\right) +2\theta_{n+m-2}\left({x}\right)+\ldots +2\theta_{|n-m|+2}\left({x}\right) +\left(1-\delta_{nm}\right)\theta_{|n-m|}\left({x}\right)\ . \end{eqnarray} The quantum numbers $J_j^{(n)}$ and $I_\alpha$ in (\ref{zaehl}) uniquely determine a particular eigenstate of the system. The asymptotic behaviour of the functions (\ref{zz}) determine their possible values. This allows to introduce densities $\rho(\nu)$ of the hole rapidities, $\sigma_n(\lambda)$ of the $\lambda$-strings and the corresponding hole densities $\tilde\rho(\nu)$ and $\tilde\sigma_n(\lambda)$ with \begin{equation} \sigma_n(x) +\tilde\sigma_n(x) = {\partial\over\partial x} z_s^{(n)}(x)\ , \quad \rho(x) + \tilde\rho(x) = {\partial\over\partial x} z_c(x)\ . \end{equation} In the thermodynamic limit $L\to\infty$ with $N_h/L$ and $M_n/L$ held fixed these equations become linear integral equations \begin{eqnarray} &&\tilde\sigma_n(x) = \left(A_{n,2S}*s\right)(x) -\sum_m \left(A_{nm}*\sigma_m\right)(x) +\left(a_n*\rho\right)(x)\ , \nonumber\\ &&\rho(x) + \tilde\rho(x) = \sum_n \left(a_n*\sigma_n\right)(x)\ . \label{intd1} \end{eqnarray} Here, $\left(f*g\right)(x)$ denotes a convolution, $2\pi a_n(x) = \theta_n'(x) = 4n/(4x^2 + n^2)$, $s(x) = 1/(2\cosh\pi x)$ and \begin{equation} A_{nm}(x) = {1\over2\pi} \Xi_{nm}'(x) + \delta_{nm}\,\delta(x)\ . \end{equation} Similarly, the energy (\ref{heigS}) in the thermodynamic limit can be rewritten in terms of the densities \begin{equation} E/L = \sum_{n=1}^{\infty} \int{\rm d}x \left(\epsilon_n^{(0)}(x)+nH\right)\sigma_n(x) - \int{\rm d}x\left(\mu+{1\over2}H\right)\rho(x) \end{equation} where $\epsilon_n^{(0)}(x)=-2\pi\left(A_{n,2S}*s\right)(x)$ are the bare energies of the $\lambda$-strings. At finite temperature the equilibrium state is obtained by minimization of the free energy $F=E-TS$ by variation of $\sigma_n$ and $\rho$. Here $S$ is the combinatorical entropy \cite{yaya:69} \begin{eqnarray} S/L &=& \sum_{n=1}^\infty \int{\rm d}x\left\{ \left(\sigma_n+\tilde\sigma_n\right) \ln\left(\sigma_n+\tilde\sigma_n\right) -\sigma_n\ln\sigma_n -\tilde\sigma_n\ln\tilde\sigma_n\right\} \nonumber\\ &&\qquad+ \int{\rm d}x \left\{ \left(\rho+\tilde\rho\right) \ln\left(\rho+\tilde\rho\right) -\rho\ln\rho -\tilde\rho\ln\tilde\rho\right\}\ . \end{eqnarray} As a result we obtain the thermodynamic Bethe ansatz (TBA) equations for the energies $\epsilon_n = T\ln(\tilde\sigma_n/\sigma_n)$ of $\lambda$-strings and $\kappa = T\ln(\tilde\rho/\rho)$ for the hole rapidities \begin{eqnarray} && \epsilon_n(x) -{T\over2\pi} \sum_m \Xi_{nm}'*\ln\left[1+{\rm e}^{-\epsilon_{m}/T}\right](x) + T a_n*\ln\left[1+{\rm e}^{-\kappa/T}\right](x) = \epsilon_n^{(0)}(x) +nH \nonumber\\ && \kappa(x) + T \sum_m a_m*\ln\left[1+{\rm e}^{-\epsilon_m/T}\right](x) = -\left(\mu+{1\over2}H\right) \label{tba0} \end{eqnarray} An alternative form of these equations can be obtained by using the identity $\sum_k \left(C_{nk}*A_{km}\right)(x) = \delta_{nm}\, \delta(x)$ with \begin{equation} C_{nm}(x) =\delta_{nm} \delta(x) - \left(\delta_{n+1,m}+\delta_{n-1,m}\right)s(x)\ . \end{equation} This allows to rewrite the integral eqs.~(\ref{intd1}) as \begin{eqnarray} \delta_{n,2S}\ s(x) &=& \sigma_n(x) + \left(C_{nm}*\tilde\sigma_m\right)(x) - \delta_{n,1}\left(s*\rho\right)(x) \nonumber\\ \left(a_{2S}*s\right)(x) &=& \tilde\rho(x) + \left([1 + a_2]^{-1}*\rho\right)(x) + \left(s*\tilde\sigma_1\right)(x)\ . \label{intd} \end{eqnarray} Similarly, we find for the energy of the corresponding state \begin{eqnarray} E/L &=& E_0^{(S)}/L -\int{\rm d}x \left[2\pi(a_{2S}*s)+\mu\right]\rho(x) +\int{\rm d} x 2\pi s(x) \tilde\sigma_{2S}(x) \\ && - \lim_{n\rightarrow\infty}Hn\int {\rm d} x \tilde\sigma_n(x)\ \nonumber \end{eqnarray} where $E_0^{(S)}$ is the ground state energy of the spin-$S$ Takhtajan--Babujian chain in a vanishing magnetic field \cite{babu:83} \begin{equation} E_0^{(S)} = \left\{ \begin{array}{ll} -\sum_{k=1}^S {2\over 2k-1} & \hbox{for~integer}\,S\\[4pt] -2\ln2-\sum_{k=1}^{S-1/2} {1\over k} & \hbox{for~half-integer}\,S \end{array} \right. \end{equation} Finally, an equivalent form of the TBA equations (\ref{tba0}) is \begin{eqnarray} \epsilon_n(x) &=& T\left(s*\ln\left[1 + {\rm e}^{\epsilon_{n -1}/T}\right] \left[1 + {\rm e}^{\epsilon_{n +1}/T}\right]\right)(x) \nonumber\\ &&\qquad - 2\pi \delta_{n,2S}\,s(x) - \delta_{n,1}T \left(s*\ln\left[1 + {\rm e}^{- \kappa/T}\right]\right)(x)\ , \label{tba1} \end{eqnarray} subject to the condition $\lim_{n\to\infty}(\epsilon_n/n) = H$ and \begin{equation} -[ 2\pi a_{2S}*s(x) + \mu] - T\left(s*\ln\left[1 + {\rm e}^{\epsilon_{1}/T}\right]\right)(x) = \kappa(x) + T\left(R*\ln\left[1 + {\rm e}^{- \kappa/T}\right]\right)(x)\ \label{kappa} \end{equation} where $R = a_2*(1 + a_2)^{-1}$. In terms of the solutions to these equations the free energy reads \begin{equation} F/L = E_0^{(S)}/L -T\int{\rm d}x s(x) \ln\left[1 + {\rm e}^{\epsilon_{2S}(x)/T}\right] -T\int{\rm d}x (a_{2S}*s)(x)\, \ln\left[1 + {\rm e}^{-\kappa(x)/T}\right]\ . \label{FreeE} \end{equation} \section{Zero temperature phases in a magnetic field} \label{sec:phasH} In the limit $T\to0$ the TBA eqs.\ (\ref{tba0}) become linear integral equations. As a consequence of (\ref{tba1}) only $\epsilon_1(x)$, $\epsilon_{2S}(x)$ and $\kappa(x)$ can take negative values for certain $x$. Hence we have to solve three coupled integral equations for these quantities which in turn determine all other dressed energies \begin{eqnarray} && \epsilon_n(x) + {1\over2\pi}\left\{ \Xi_{n1}'*\epsilon_1^{(-)} + \Xi_{n,2S}'*\epsilon_{2S}^{(-)}\right\}(x) - a_n*\kappa^{(-)}(x) = \epsilon_n^{(0)}(x) +nH \nonumber\\ && \kappa(x) - \left\{ a_1*\epsilon_1^{(-)} + a_{2S}*\epsilon_{2S}^{(-)}\right\}(x) = -\left(\mu+{1\over2}H\right)\ , \label{dressE} \end{eqnarray} where $f^{(\pm)}(x) = \theta(\pm f(x)) f(x)$. To discuss the solutions of these equations further we have to distinguish various cases: \subsection{$\mu>H/2$} In this regime we have $\kappa(x)<0$ and $\epsilon_1(x)<0$ for all $x$. This allows to express these functions in terms of the remaining unknown function $\epsilon_{2S}^{-}(x)$. From (\ref{dressE}) we find \begin{eqnarray} && \kappa(x) = -2\mu+\left(a_1*\epsilon_1^{(0)}\right)(x) \nonumber\\ && \epsilon_1(x) = \epsilon_1^{(0)}(x) - \left(\mu-{H\over2}\right) -\left(a_{2S-1}*\epsilon_{2S}^{(-)}\right)(x) \nonumber\\ && \epsilon_{n>1}(x) + {1\over2\pi}\left( \Xi_{n-1,2S-1}'*\epsilon_{2S}^{(-)}\right)(x) = -2\pi\left(A_{n-1,2S-1}*s\right)(x)+(n-1)H \nonumber \end{eqnarray} The last set of equations can be identified with the integral equations for the dressed energies of the spin-$S-1/2$ Takhtajan--Babujian model, hence this regime corresponds to the completely doped case (i.e.\ $N_h=L$ holes). For magnetic field $H>H^{(S-1/2)}$ with \begin{equation} H^{(\sigma)}>{2\over\sigma}\sum_{k=1}^{2\sigma}{1\over2k-1} \label{Hsat1} \end{equation} the system is in a ferromagnetically saturated state with maximal magnetization $M^z=L(S-1/2)$. \subsection{$-H/2<\mu<H/2$} Here we find from (\ref{dressE}) that $\kappa(x)\equiv \kappa^{(-)}(x)<0$ for all $x$ while $\epsilon_1(x)$ can take non-negative values. Eliminating $\kappa(x)$ from the integral equations for $\epsilon_1$ and $\epsilon_{2S}$ we obtain \begin{eqnarray} && \epsilon_1(x) + \left\{a_{2S-1}*\epsilon_{2S}^{(-)}\right\}(x) = \epsilon_1^{(0)}(x) - \mu+{1\over2}H\,, \nonumber\\ && \epsilon_{2S}(x) + \left\{a_{2S-1}*\epsilon_{1}^{(-)} + 2\sum_{k=1}^{2S-1}a_{2k}*\epsilon_{2S}^{(-)} \right\}(x) = \epsilon_{2S}^{(0)}(x) - \mu+{4S-1\over2}H\, . \nonumber \end{eqnarray} In this regime we find $\epsilon_{1}>0$ and $\epsilon_{2S}>0$ for \begin{equation} \mu< \min\left\{{4S-1\over2}H -4\sum_{k=1}^{2S}{1\over2k-1}\, ,\, {1\over2}H-{2\over S}\right\}\ . \label{ferro1} \end{equation} Positive dressed energies for the $\lambda$-strings imply $M_n=0$ and from (\ref{intd1}) we find that $N_h=0$ in this region. Hence for $T\to 0$ (\ref{ferro1}) belongs to the ferromagnetically saturated phase of the undoped system, namely the spin-$S$ Takhtajan-Babujian model. This phase exists for magnetic fields $H>H^{(S)}$. Increasing the hole chemical potential to values $\mu>H/2-2/S$ holes are added but the ground state continues to be fully polarized: For the dressed energies this corresponds to $\epsilon_{2S}>0$ and while the real spin rapidities fill all states with negative $\epsilon_1(x)$. As a consequence of the free fermionic nature of this state these dressed energies can be expressed in terms of their free values \begin{equation} \epsilon_{1}(x) = \epsilon_1^{(0)}(x) -\left(\mu-{H\over2}\right)\ . \end{equation} The lower boundary of this phase in the $\mu$--$H$ plane is given by the condition $\min_x\left\{\epsilon_{2S}(x)\right\}=0$. For magnetic fields $(4S-1/2)H<\mu+4\sum_{k=1}^{2S}1/(2k-1)$ the ground state is a filled sea of $\lambda$-strings of length $2S$ with negative energy \begin{equation} \epsilon_{2S}(x) + \left\{\sum_{k=1}^{2S-1} a_{2k}*\epsilon_{2S}^{(-)}\right\}(x) = \epsilon_{2S}^{(0)}(x) + \left(2S-{1\over2}\right)H -\mu\ . \label{phaseX} \end{equation} The other dressed energies $\kappa(x)<0$ and $\epsilon_{n\ne2S}(x)>0$ can be expressed in terms of the solution of this equation. In this region of parameters the system has a finite concentration of holes \emph{and} overturned spins. Although one might na\"{\i}vely expect two types of massless excitations in such situation only one such branch with dispersion (\ref{phaseX}) is realized in this system which turns out to describe the charge excitations. Hence spin excitations are gapped in this regime \cite{frso:pp}. \subsection{$\mu<-H/2$} \label{ssec:phas3} Again we find several phases that can be characterized by the configurations of strings present in the ground state: For $H>H^{(S)}$ all dressed energies are positive corresponding to completely polarized undoped state. For smaller magnetic fields $\epsilon_{2S}(x)$ takes negative values in some interval to be determined from \begin{equation} \epsilon_n(x) + {1\over2\pi}\left\{ \Xi_{n,2S}'*\epsilon_{2S}^{(-)}\right\}(x) = \epsilon_n^{(0)}(x) +nH\ . \end{equation} These are the Bethe ansatz equations of the spin-$S$ Takhtajan--Babujian chain. This phase becomes unstable against hole creation for chemical potentials \begin{equation} \mu>\int {\rm d}a_{2S}(x)\epsilon_{2S}^{(-)}(x) -{1\over2}H \to \psi\left({2S+1\over4}\right) -\psi\left({2S+3\over4}\right) \mathrm{~for~}H=0 \end{equation} ($\psi(x)$ is the digamma function). Beyond this line the ground state is built from a filled sea of $\lambda$-strings with energies $\epsilon_{2S}<0$ and another sea of charge rapidities with energies $\kappa(x)<0$. Increasing the chemical potential further negative energy solutions for the real spin rapidities $\epsilon_1$ appear giving rise to a third condensate of Bethe rapidities. From the cases considered above we obtain the qualitative zero temperature phase diagram of the doped spin-$S$ system in the $\mu$--$H$ plane presented in Fig.~\ref{fig:phasemu}(a). Using Eqs.~(\ref{intd1}) the corresponding phase boundaries can be given as a function of the hole concentration $x=N_h/L$. For $S=1$ this is shown in Fig.~\ref{fig:phasemu}(b). \section{Low-temperature thermodynamics at $H=0$} \label{sec:phasT} Further simplification is possible in the case of finite doping in a vanishing magnetic field which corresponds to chemical potentials $\mu \in \left[\psi((2S+1)/4)-\psi((2S+3)/4),0\right]$. Furthermore, $\epsilon_1(x)<0$, $\epsilon_{2S}(x)<0$ for all $x$ while $\kappa(x)<0$ for $|x| < Q$ in this regime. All other dressed energies vanish in this limit. Eliminating the $\epsilon_n$ from the equation for the energy of the holes we obtain \begin{equation} -[ 2\pi a_{2S}*s(x) + \mu] = \kappa(x) -\int_{-Q}^{Q}{\rm d}y R(x-y)\kappa(y)\ \label{kappa0} \end{equation} where the Fermi point $Q$ is determined by the condition $\kappa(\pm Q)=0$. Similarly, the density of hole rapidities $\rho_0(x)$ in this regime is given by \begin{eqnarray} \rho_0(x) - \int_{-Q}^Q {\rm d} yR(x - y)\rho_0(y) = a_{2S}*s(x)\ . \end{eqnarray} Excitations with charge rapidities near $\pm Q$ are massless. The velocity of this charge mode can be obtained from the dispersion (\ref{kappa0}) \begin{equation} v = \frac{1}{2\pi\rho_0(Q)} \left.\frac{\partial\kappa_0}{\partial x}\right|_{x = Q}\ . \label{v} \end{equation} Similarly, one has massless excitations near $x=\pm\infty$ in the magnetic sector with energies $\epsilon_{1}(x)$ and $\epsilon_{2S}(x)$ with velocities \begin{equation} v_{2S} = \lim_{x\to\infty} {\epsilon_{2S}'(x)\over2\pi\sigma_{2S}(x)} \equiv \pi\ , \quad v_{1} = \lim_{x\to\infty} {\epsilon_{1}'(x)\over2\pi\sigma_1(x)} = -{1\over2}\, \frac{\int_{-Q}^Q{\rm d} y {\rm e}^{\pi y}\kappa_0(y)}{ \int_{-Q}^Q{\rm d} y{\rm e}^{\pi y}\rho_0(y)}\ . \label{veloS} \end{equation} As a consequence of the behaviour of the dressed energies as $H\to0$ we can replace $\kappa$ in Eq.~(\ref{tba1}) by its zero temperature value $\kappa_0(x)$ and the driving terms by their asymptotics to obtain the leading low temperature behaviour. As a result we get \begin{equation} \epsilon_n(x) = Ts*\ln[1 + {\rm e}^{\epsilon_{n -1}(x)/T}][1 + {\rm e}^{\epsilon_{n +1}(x)/T}] - 2\pi\delta_{n,2S}\,{\rm e}^{- \pi|x|} - 2\pi A\delta_{n,1}{\rm e}^{- \pi|x|} \label{tba2} \end{equation} where $2\pi A = - \int_{-Q}^Q{\rm d} y {\rm e}^{\pi y}\kappa_0(y)$. To move further we have to separate the contributions to the free energy stemming from the charge-sector from those due to the $\epsilon_n$. Considering low temperatures again the leading contributions to $\kappa$ come from the vicinity of the Fermi wave vectors $\pm Q$. In this region one can safely neglect contributions to Eq.~(\ref{kappa}) from $\epsilon_1$ and rewrite it as \begin{eqnarray} && -[2\pi a_{2S}*s(x) + \mu] - TR*\ln[1 + {\rm e}^{- |\kappa(x)|/T}] \nonumber\\ &&\qquad = \kappa(x) - \int_{-Q}^Q {\rm d} yR(x - y)\kappa(y) \end{eqnarray} where $Q$ is determined by the condition $\kappa(\pm Q) = 0$. Using the procedure introduced by Takahashi \cite{taka:71b}, we can rewrite the free energy (\ref{FreeE}) as $F/L = E_0^{(S)}/L + f_c + f_s$ where \begin{eqnarray} f_c &=& - T\int{\rm d}x \rho_0(x) \ln\left[1 + {\rm e}^{- |\kappa_0(x)|/T}\right] \approx- \pi T^2/6v\ , \label{Free}\\ f_s &=& - T\int{\rm d}x s(x) \ln\left[1 + {\rm e}^{\epsilon_{2S}(x)/T}\right] - T\int{\rm d}x (s*\rho_0)(x) \ln\left[1 + {\rm e}^{\epsilon_1(x)/T}\right]\ . \label{free2} \end{eqnarray} Now the thermodynamics is described by Eq.~(\ref{Free}) representing a scalar bosonic mode (the charge sector) and by Eqs.~(\ref{tba2}) and (\ref{free2}) for the spin sector. At low temperatures the spin contribution $f_s$ is dominated by contributions from the regions $v_{2S}\exp(- \pi |x|) \sim T$ where $|\epsilon_{2S}| \sim T$ and the second one by the regions $v_1\exp(- \pi |x|) \sim T$. The leading temperature dependence of $f_s$ at low $T$ can be obtained by rewriting (\ref{tba2}) for large positive $x$ as \begin{equation} \varphi_n(x) = s*\ln[1 + {\rm e}^{\varphi_{n -1}}][1 + {\rm e}^{\varphi_{n+1}}] - \delta_{n,2S}\,{\rm e}^{- \pi x} - A\delta_{n,1}{\rm e}^{- \pi x} \label{tba22} \end{equation} in terms of the $T$-independent functions \[ \varphi_n(x) = {1\over T} \epsilon_n\left(x-{1\over\pi}\ln{T\over2\pi}\right)\ . \] In the low-$T$ limit we can replace $s(x)$ and $s*\rho_0(x)$ in (\ref{free2}) by their asymptotics to obtain the free energy \begin{equation} f_s \simeq - {\pi T^2\over6} \left( {c_{2S}\over v_{2S}} + { c_1 \over v_1} \right)\ . \label{free22} \end{equation} Such an expression is characteristic of a system two with massless excitations with velocities (\ref{veloS}). In cases where these excitations can be characterized by different \emph{observable} quantum numbers the coefficients $c_i$ are the central charges of the underlying Virasoro-algebra thus determining the universality class of the system. In this case they are given in terms of the solutions of (\ref{tba22}) by \begin{equation} c_{2S} = {6\over\pi} \int{\rm d}x\, {\rm e}^{-\pi x} \ln\left[1 + {\rm e}^{\varphi_{2S}(x)}\right]\ , \qquad c_1 = {6\over\pi}\int{\rm d}x\, (A {\rm e}^{-\pi x}) \ln\left[1 + {\rm e}^{\varphi_1(x)}\right]\ . \label{charges} \end{equation} Using standard methods \cite{babu:83,klme:90,zamo:91} for the analysis of the TBA equations we find that their sum can be written as \begin{equation} c_{2S}+c_{1} = {6\over\pi^2} \sum_n \left[ {\cal L}\left({{\rm e}^{\varphi_n(x)} \over1+{\rm e}^{\varphi_n(x)}}\right) \right]_{x=-\infty}^\infty \label{csum} \end{equation} where ${\cal L}(x)$ is Rogers dilogarithm function \[ {\cal L}(x) = -{1\over2}\int_0^x {\rm d}t\, \left[ {\ln t\over 1-t} + {\ln(1-t)\over t}\right]\ . \] Hence, the $c_{2S}+c_1$ is completely determined by the asymptotic behaviour of the solutions of (\ref{tba22}) as $x\to\pm\infty$: \begin{eqnarray} \lim_{x\to\infty}\varphi_n(x) &=& \ln\left((n+1)^2-1\right)\ , \nonumber\\ \lim_{x\to-\infty}\varphi_n(x) &=& \left\{\begin{array}{ll} \ln\left((n-2S+1)^2-1\right) & {\rm for~}n>2S \\ \ln\left({\sin^2(\pi n/2S+1)\over\sin^2(\pi/2S+1)} -1\right) & {\rm for~}1<n<2S\\ -\infty & {\rm for~} n=1,\,2S \end{array} \right.\ , \nonumber \end{eqnarray} giving \begin{equation} c_{2S}+c_1 = 2 {4S-1\over2S+1}\ \label{SumC} \end{equation} independent of the doping (i.e.\ $A$). The individual values of the $c_i$ are easily calculated for small and large doping corresponding to $A\to0$ and $A\to\infty$, respectively. In these cases the regions contributing to the integrals (\ref{charges}) are well separated and the functions $\varphi_n(x)$ take constant values in between. For small doping ($A\ll1$) we find $\varphi_n(x)=\varphi_n^{(0)}$ for $\ln A\ll \pi x\ll 0$ with \begin{equation} \varphi_n^{(0)} = \left\{\begin{array}{ll} \ln\left((n-2S+1)^2-1\right) & {\rm for~}n>2S \\ -\infty & {\rm for~} n=2S\\ \ln\left({\sin^2(\pi n/2(S+1))\over\sin^2(\pi/2(S+1))} -1\right) & {\rm for~}1\le n<2S \end{array} \right.\, . \end{equation} Hence, near $x\approx0$ they behave as in the undoped system giving the central charge $3S/(S+1)$ of the $SU(2)_{2S}$ WZNW model. In the region around $x\approx \ln A$ the $\varphi_{n<2S}$ are solutions of the {\em finite} set of TBA of the minimal unitary model ${\cal M}_{p}$ \cite{zamo:91} with central charge $c_1=1-6/(p(p+1))$ where $p=2S+1$ (this is the Ising model for $S=1$ (see \cite{fpt:98}), tricitical Ising model for $S=3/2$, three-state Potts model for $S=2$, tricritical three state Potts model for $S=5/2$ and so forth). Putting everything together we find the leading contribution to the spin part (\ref{free2}) to the free energy at small doping \begin{equation} f_s = -{\pi T^2\over 6v_{2S}}\, {3S\over S+1} -{\pi T^2\over 6v_1}\left\{1-{3\over{(S+1)(2S+1)}}\right\}\, . \label{fs_0} \end{equation} Proceeding in an analogeous way in the limit of large doping ($A\gg1$) corresponding to a spin-$(S-1/2)$ chain doped with spin-$S$ carriers we find \begin{equation} \varphi_n^{(0)} = \left\{\begin{array}{ll} \ln\left(n^2-1\right) & {\rm for~}n>1 \\ -\infty & {\rm for~} n=1\\ \end{array} \right.\, \end{equation} for $0\ll \pi x\ll \ln A$. In this limit the low temperature contributions to $f_s$ can be written as the sum of a $SU(2)_1$ and a $SU(2)_{2S-1}$ WZNW model, the latter being the well known continuum limit of the pure spin-$(S-1/2)$ Takhtajan-Babujian model: \begin{equation} f_s = -{\pi T^2\over 6v_{2S}}\, {6S-3\over 2S+1} -{\pi T^2\over 6v_1}\ . \label{fs_1} \end{equation} For finite values of $A$ the coefficients $c_{2S}$ and $c_1$ in (\ref{free22}) have to be determined numerically. They are found to interpolate smoothly between their limiting values in (\ref{fs_0}) and (\ref{fs_1}). For $S=1$ and $S=3$ their doping dependence is shown in Figure~\ref{fig:ccx}. \section{Summary and Conclusion} To summarize, we have introduced a class of integrable models describing a magnetic system which upon doping interpolates between the integrable spin-$S$ and $S-1/2$ Takhtajan-Babujian chains. These models arise when considering vertex models invariant under the action of the graded Lie algebra $gl(2|1)$ with the local quantum spaces carrying the `atypical' higher-spin representations $\left[S\right]_+$. Their solution by means of the algrebraic Bethe Ansatz allows for a detailed study of their low temperature phase diagram. The spectrum of low lying excitations is described in terms of the dressed energies satisfying the TBA equations (\ref{tba1}) and (\ref{kappa}). Without an external magnetic field the critical degrees of freedom separate into charge and magnetic modes as is well known in the Tomonaga-Luttinger liquid models for one-dimensional correlated electrons (see e.g.\ Refs.~\cite{frko:90,kaya:91}). Different from these models, however, one finds \emph{two} branches of low lying modes in the magnetic sector which at small (large) doping can be identified with higher level $SU(2)_k$ WZW models and a minimal model (free boson). The WZW models have to be present in order to reproduce the well understood critical behaviour of the undoped and completely doped limiting cases. The second gapless magnetic mode, however, is quite peculiar: its appearence in the low energy of the undoped system is crucial to allow for the smooth crossover between the limiting cases (\ref{fs_0}) and (\ref{fs_1}) subject to the constraint $c_{2S}+c_1={\rm const}$. The low-$T$ behaviour of the $S=1$ integrable model has motivated the proposition of an effective field theory of four (real) Majorana fermions as a possible starting point for studies of perturbations around the integrable model \cite{fpt:98}. While free field representations could be used for the constituents of the \emph{undoped} model, interaction terms between the two sectors had to be introduced to reproduce the change of the coefficients $c_{2S}$ and $c_1$ with the hole concentration observed in the exact solution. The possible form of this interaction term is constrained by the $SU(2)$-symmetry of the model without a magnetic field. A similar construction of an $SU(2)$-invariant effective low energy field theory for the $S>1$ models introduced here is possible by using the fact that the minimal models can be obtained within a GKO coset construction applied to \cite{gko:85,gko:86} \begin{equation} {SU(2)_{2S-1}\otimes SU(2)_{1} \over SU(2)_{2S}}\ . \label{GKO} \end{equation} In fact, the observed change in the conformal weights attributed to the magnetic modes between the limiting cases of the undoped and the completely doped system appear to be just a `adiabatic' realization of this construction \begin{equation} SU(2)_{2S} \otimes {\cal M}_{2S+1} \longrightarrow SU(2)_{2S-1}\otimes SU(2)_{1}\ . \end{equation} On the other hand, taking the limit $H\to 0$ starting from the phase discussed in Section~\ref{ssec:phas3} one may obtain a different field theoretical description of the $SU(2)$-symmetric phase: There the critical degrees of freedom can be described in terms of two free bosons each contributing $c=1$ to the sum $c_{2S}+c_1$. For $S=1$ this should give a complete description of the massless magnetic modes. It is likely that the apparent difference between the $H\to0$ limit and the $H=0$ model can be understood as a rotation in the space of the effective fields (note that no \emph{physical} field couples to one of the magnetic modes alone) \cite{fab:pc}. For $S>1$ one has $c_{2S}+c_1>2$ from (\ref{SumC}). Here, the difference to the finite field critical properties is similar to the one observed in the Takhtajan-Babujian models \cite{tsve:88}: it is due to the appearence of gap for parafermionic degrees of freedom in the critical theory for any non-zero magnetic field. \section*{Acknowledgements} I am grateful to A.~M.\ Tsvelik and F.~H.~L.\ E\ss{}ler for important discussions. This work has been supported by the Deutsche Forschungsgemeinschaft under Grant No.\ Fr~737/2--3. \newpage
\section{Introduction} The electromagnetic form factors of the nucleon and its excitations (baryon resonances) provide a powerful tool to investigate the structure of the nucleon \cite{Baryons}. These form factors can be measured in electroproduction as a function of the four-momentum squared $q^2=-Q^2$ of the virtual photon. The elastic form factors of the nucleon have mostly been determined with the Rosenbluth separation technique. However, for values of $Q^2$ above a few (GeV/c)$^2$ it becomes increasingly difficult to extract the electric form factor $G_E$ because the measured cross section is dominated by the magnetic form factor $G_M$. For neutron form factors there is the additional problem of the lack of a free neutron target, and the (small) relative size of the electric versus the magnetic form factors for small values of $Q^2$. With the advent of high duty cycle electron facilities such as ELSA, MAMI, and Jefferson Laboratory far more accurate determinations of the form factors become feasible (ratio method, polarization variables), especially for the neutron form factors (see {\em e.g.} \cite{gendat,gmndat2}). Recent experiments on electroproduction and eta-photoproduction have yielded valuable new information on transition form factors, in particular on the helicity amplitudes of the N(1520)$D_{13}$ and N(1535)$S_{11}$ resonances \cite{Nimai1,Nimai2,Armstrong}. The challenge for theoretical calculations is to provide a simultaneous description of both the four elastic form factors and the transition form factors, even if it is only in a phenomenological approach. In this contribution we present such an analysis in the framework of a recently introduced algebraic model of the nucleon \cite{BIL}. \section{Algebraic model} The algebraic approach provides a unified treatment of various constituent quark models \cite{BIL}, such as harmonic oscillator quark models and collective models. In this paper we employ a collective model of the nucleon in which baryon resonances are interpreted as vibrational and rotational excitations of an oblate top. There are two fundamental vibrations: a breathing mode and a two-dimensional vibrational mode, which are associated with the $N(1440)P_{11}$ Roper resonance and the $N(1710)P_{11}$ resonance, respectively. The negative parity resonances of the second resonance region are interpreted as rotational excitations. Since each vibrational mode has its own characteristic frequency, there is no problem with the mass of the Roper resonance relative to that of the negative parity resonances. The electromagnetic form factors are obtained by folding with a distribution of the charge and magnetization over the entire volume of the baryon \cite{BIL}. All calculations are carried out in the Breit frame. \section{Elastic form factors} In \cite{emff} we studied the elastic electromagnetic form factors of the nucleon. These calculations include anomalous magnetic moments for the proton and the neutron, as well as a flavor dependent distribution functions of the charge and magnetization. Supposedly, the anomalous magnetic moments and the flavor dependence arise as effective parameters, since the coupling to the meson cloud surrounding the nucleon was not included explicitly. According to \cite{emff} the electric and magnetic form factors of the nucleon, when folded with a distribution of the charge and magnetization, can be expressed in terms of a common intrinsic dipole form factor \begin{eqnarray} G_E^p(Q^2) &=& g(Q^2) ~, \nonumber\\ G_E^n(Q^2) &=& 0 ~, \nonumber\\ G_M^p(Q^2) &=& g(Q^2) ~, \nonumber\\ G_M^n(Q^2) &=& -2g(Q^2)/3 ~, \label{Sachs} \end{eqnarray} with \begin{eqnarray} g(Q^2) &=& \frac{1}{(1+\gamma Q^2)^2} ~. \end{eqnarray} Note that the form factors of Eq.~(\ref{Sachs}) do not contain anomalous magnetic moments nor involve flavor dependent distribution functions. In order to study the coupling to the meson cloud we express the electric and magnetic form factors in terms of their isoscalar ($S$) and isovector ($V$) components \begin{eqnarray} G_{E}^S(Q^2) &=& G_{E}^p(Q^2) + G_{E}^n(Q^2) \;=\; g(Q^2) ~, \nonumber\\ G_{E}^V(Q^2) &=& G_{E}^p(Q^2) - G_{E}^n(Q^2) \;=\; g(Q^2) ~, \nonumber\\ G_{M}^S(Q^2) &=& G_{M}^p(Q^2) + G_{M}^n(Q^2) \;=\; g(Q^2)/3 ~, \nonumber\\ G_{M}^V(Q^2) &=& G_{M}^p(Q^2) - G_{M}^n(Q^2) \;=\; 5g(Q^2)/3 ~. \end{eqnarray} \subsection{Meson cloud couplings} The effects of the meson cloud surrounding the nucleon are taken into account by including the coupling to the isoscalar vector mesons $\omega$ and $\phi$ and the isovector vector meson $\rho$. These contributions are studied phenomenologically by parametrizing the isoscalar and isovector components of the form factors as \begin{eqnarray} G_E^S(Q^2) &=& g(Q^2) \left[ \alpha^S + \alpha_{\omega} \, \frac{m_{\omega}^2}{m_{\omega}^2+Q^2} + \alpha_{\phi} \, \frac{m_{\phi}^2}{m_{\phi}^2+Q^2} \right] ~, \nonumber\\ G_E^V(Q^2) &=& g(Q^2) \left[ \alpha^V + \alpha_{\rho} \, \frac{m_{\rho}^2}{m_{\rho}^2+Q^2} \right] ~, \nonumber\\ G_M^S(Q^2) &=& \frac{1}{3} g(Q^2) \left[ \beta^S + \beta_{\omega} \, \frac{m_{\omega}^2}{m_{\omega}^2+Q^2} + \beta_{\phi} \, \frac{m_{\phi}^2}{m_{\phi}^2+Q^2} \right] ~, \nonumber\\ G_M^V(Q^2) &=& \frac{5}{3} g(Q^2) \left[ \beta^V + \beta_{\rho} \, \frac{m_{\rho}^2}{m_{\rho}^2+Q^2} \right] ~. \label{SV} \end{eqnarray} The large width of the $\rho$ meson ($\Gamma_{\rho}=151$ MeV) is taken into account by making the replacement \cite{IJL} \begin{eqnarray} \frac{m_{\rho}^2}{m_{\rho}^2+Q^2} &\rightarrow& \frac{m_{\rho}^2 + 8 \Gamma_{\rho} m_{\pi}/\pi} {m_{\rho}^2+Q^2 + (4m_{\pi}^2+Q^2) \Gamma_{\rho} \alpha(Q^2)/m_{\pi}} ~, \end{eqnarray} with \begin{eqnarray} \alpha(Q^2) &=& \frac{2}{\pi} \left[ \frac{4m_{\pi}^2+Q^2}{Q^2} \right]^{1/2} \ln \left( \frac{\sqrt{4m_{\pi}^2+Q^2} + \sqrt{Q^2}}{2m_{\pi}} \right) ~. \end{eqnarray} The coefficients $\alpha^{S/V}$ and $\beta^{S/V}$ in Eq.~(\ref{SV}) are determined by the electric charges and the magnetic moments of the nucleon, respectively \begin{eqnarray} \alpha^S &=& 1-\alpha_{\omega}-\alpha_{\phi} ~, \nonumber\\ \alpha^V &=& 1-\alpha_{\rho} ~, \nonumber\\ \beta^S &=& 3(\mu_p + \mu_n) - \beta_{\omega}-\beta_{\phi} ~, \nonumber\\ \beta^V &=& \frac{3}{5} (\mu_p - \mu_n) - \beta_{\rho} ~. \end{eqnarray} For small values of the momentum transfer the form factors are dominated by the meson dynamics and reduce to a monopole form, whereas for large values the modification of dimensional counting laws from perturbative QCD is taken into account by scaling $Q^2$ with the strong coupling constant \cite{GK,Speth} \begin{eqnarray} Q^2 &\rightarrow& Q^2 \frac{\alpha_s(0)}{\alpha_s(Q^2)} \;=\; Q^2 \frac{\ln [(Q^2 + \Lambda_{asy}^2)/\Lambda_{QCD}^2]} {\ln [\Lambda_{asy}^2/\Lambda_{QCD}^2]} ~. \end{eqnarray} \subsection{Results} In Figs.~\ref{gepfd}--\ref{gmnfd} we show a compilation of the most recent data on the electromagnetic form factors of the nucleon. Most of the data points have been obtained by a separation technique which is based on the Rosenbluth formula for the cross section \cite{Rosenbluth} \begin{eqnarray} \frac{d \sigma}{d \Omega} &=& \left( \frac{d \sigma}{d \Omega} \right)_{Mott} \left[ \frac{G_E^2(Q^2) + \tau G_M^2(Q^2)}{1+\tau} +2\tau G_M^2(Q^2) \tan^2 \frac{\theta}{2} \right] ~, \label{sigma} \end{eqnarray} with $\tau=Q^2/4M^2$. Although in principle the electric and magnetic form factors can be separated by varying the scattering angle $\theta$, in practice this separation technique is limited to small values of $Q^2$ only. For large values of $Q^2$ the measured cross section is dominated by the magnetic form factor, and hence it becomes increasingly difficult to extract the electric form factor. For the neutron form factors there is the additional problem of the lack of a free neutron target, and the (small) relative size of the electric versus the magnetic form factors for small values of $Q^2$. In order to overcome these problems to extract the form factors many times the assumption of form factor scaling (which holds for small values of $Q^2$) is made \begin{eqnarray} G_E^p(Q^2) \;=\; \frac{G_M^p(Q^2)}{\mu_p} \;=\; \frac{G_M^n(Q^2)}{\mu_n} ~. \label{scaling} \end{eqnarray} As a result, especially for the neutron form factors the scattering of data points is at times even larger than the uncertainties quoted by the authors. With the advent of high duty cycle electron facilities such as {\em e.g.} ELSA, MAMI and Jefferson Laboratory far more accurate determinations of the form factors become feasible, either using coincidence experiments (ratio method) or polarization variables \cite{gendat,gmndat2}. For example, the use of polarization variables makes it possible to extract the ratio of the electric and magnetic form factors from the data. In the present calculation the coefficients $\alpha_M$ and $\beta_M$ (with $M=\rho$, $\omega$, $\phi$), the scale parameter $\gamma$ in the dipole form factor $g(Q^2)$ and the $\Lambda$'s are determined in a simultaneous fit to all four electromagnetic form factors of the nucleon and the proton and neutron charge radii. The values of the fitted parameters are given in Table~\ref{parameters}. As usual, the form factors are scaled by the standard dipole fit $F_D=1/(1+Q^2/0.71)^2$. The oscillations around the dipole values are attributed to the meson cloud couplings. The electric form factors, as well as the proton and neutron charge radii (the slope of $G_E^p$ and $G_E^n$ in the origin) are reproduced well. The electric form factor of the neutron is the least known. Unlike for the proton, the Rosenbluth separation of $G_E^n$ from $G_M^n$ for a neutron target is difficult for all values of $Q^2$: for small $Q^2$ because of the small size of $G_E^n$ compared to $G_M^n$, and for large $Q^2$ because the magnetic component dominates both the angular dependent and angular independent term in the cross section. For this reason we have included in Fig.~\ref{gen} only the results obtained from polarization variables \cite{gendat}. The new data points for $G_E^n$ are significantly larger than those of the Platchkov compilation \cite{Platchkov}, which in turn leads to a reduction of the cross-over point in the neutron charge distribution from 0.9 fm to 0.7 fm \cite{Drechsel}. In the Breit frame, the proton and neutron charge distributions are given by the Fourier transforms of the respective electric form factors \begin{eqnarray} \rho_{p/n}(r) = \frac{1}{(2\pi)^3} \int d\vec{q} \, G_E^{p/n}(q) \, \mbox{e}^{-i \vec{q} \cdot \vec{r}} ~. \end{eqnarray} In Figs.~\ref{rhop} and \ref{rhon} we show the results of our calculations. The neutron charge distribution shows a change in sign at 0.65 fm. The magnetic form factors show a slight oscillatory behavior around the dipole form. For large values of $Q^2$ both the proton and the neutron magnetic form factor show a decrease with respect to the dipole. The present calculation is in good agreement with the new measurements of $G_M^n$ at MAMI (Anklin '98 \cite{gmndat2}). In Figs.~\ref{gepgmp} and \ref{gmngmp} we study the form factor scaling relations of Eq.~(\ref{scaling}). Whereas the ratio of the proton form factors is close to one over the entire range of $Q^2$, the ratio of neutron and proton magnetic form factors shows a deviation from form factor scaling. In Figs.~\ref{snsp} and \ref{fprat} we show two other measures of deviations from form factor scaling. The ratio of neutron and proton Rosenbluth cross sections reduces under the assumption of form factor scaling and for large values of $Q^2$ to a constant \begin{eqnarray} \left. \frac{d \sigma_n}{d \Omega} \right/ \frac{d \sigma_p}{d \Omega} &=& \frac{(G_E^n)^2 + \tau (G_M^n)^2 +2\tau(1+\tau) (G_M^n)^2 \tan^2 (\theta_n/2) } {(G_E^p)^2 + \tau (G_M^p)^2 +2\tau(1+\tau) (G_M^p)^2 \tan^2 (\theta_p/2)} \nonumber\\ &\rightarrow& \frac{\tau \mu_n^2 +2\tau(1+\tau) \mu_n^2 \tan^2 (\theta_n/2)} {1 + \tau \mu_p^2 +2\tau(1+\tau) \mu_p^2 \tan^2 (\theta_p/2)} \nonumber\\ &\rightarrow& \frac{\mu_n^2 \tan^2 (\theta_n/2)} {\mu_p^2 \tan^2 (\theta_p/2)} ~. \label{sigmanp} \end{eqnarray} For equal angles ($\theta_n = \theta_p$) this ratio reduces to $\mu_n^2/\mu_p^2 = 0.47$~. The transition from the region of low momentum transfer where to good approximation the nucleon form factors satisfy form factor scaling, to the region of high momentum transfer for which the methods of perturbative QCD apply, can be studied by the ratio $Q^2 F_2^p / F_1^p$ of the Dirac ($F_1$) and Pauli ($F_2$) form factors \begin{eqnarray} F_1(Q^2) &=& \frac{G_E(Q^2) + \tau G_M(Q^2)}{1+\tau} ~, \nonumber\\ F_2(Q^2) &=& \frac{G_M(Q^2) - G_E(Q^2)}{(\mu_p-1)(1+\tau)} ~. \end{eqnarray} According to dimensional scaling laws the helicity conserving amplitude $F_1^p$ dominates the helicity-flip amplitude $F_2^p$ at high $Q^2$, and the ratio $Q^2 F_2^p / F_1^p$ goes to a constant \cite{Brodsky}. Under the assumption of form factor scaling and for large values of $Q^2$ we find \begin{eqnarray} \frac{Q^2 F_2^{p}}{F_1^{p}} &=& \frac{Q^2(G_M^{p} - G_E^{p})}{(\mu_p-1)(G_E^{p} + \tau G_M^{p})} \nonumber\\ &\rightarrow& \frac{Q^2}{1 + \tau \mu_p} \nonumber\\ &\rightarrow& \frac{4M^2}{\mu_p} \;=\; 1.26 \mbox{ (GeV/c)}^2 ~. \label{f2f1} \end{eqnarray} In Figs.~\ref{snsp} and \ref{fprat} both the data and our calculations show a saturation with increasing values of $Q^2$. A comparison between the full calculation (solid lines, labeled `present') and the assumption of form factor scaling (dashed lines, labeled `scaling') shows that for the ratio of the neutron and proton cross sections there is a large deviation from form factor scaling. For the ratio $Q^2 F_2^p/F_1^p$ the two curves coincide. For comparison we also show the results of two other calculations: the vector meson dominance model of Iachello, Jackson and Lande \cite{IJL} (dash-dotted lines, labeled `IJL'), and a hybrid model (interpolation between vector meson dominance and pQCD) by Gari and Kr\"umpelmann \cite{GK} (dash-dashed lines, labeled `GK'). In comparing the different calculations one has to keep in mind that each one was optimized with the data set of that time (1973 for \cite{IJL} and 1985 for \cite{GK}). \section{Transition form factors} In addition to the elastic form factors, there is currently much interest in the inelastic transition form factors. Recent experiments on eta-photoproduction $\gamma + N \rightarrow N^{\ast} + \eta$ have yielded valuable new information on the helicity amplitudes of the N(1520)$D_{13}$ and N(1535)$S_{11}$ resonances. In an Effective Lagrangian Approach these new experimental results were used to extract model independent ratios of photocouplings \cite{Nimai1,Nimai2} \begin{eqnarray} \mbox{N(1535)}S_{11} &: \hspace{1cm}& \frac{A^n_{1/2}}{A^p_{1/2}} \;=\; -0.84 \pm 0.15 ~. \nonumber\\ \mbox{} \nonumber\\ \mbox{N(1520)}D_{13} &: \hspace{1cm}& \frac{A^p_{3/2}}{A^p_{1/2}} \;=\; -2.5 \pm 0.2 \pm 0.4 ~. \end{eqnarray} These values are in excellent agreement with those of the collective algebraic model, $-0.81$ and $-2.53$, respectively \cite{BIL} (for the N(1535)$S_{11}$ resonance a mixing angle of $\theta = -38$ degrees was introduced). In Table~\ref{ratios} we compare these values with some model calculations. Whereas for the ratio $A^n_{1/2}/A^p_{1/2}$ of the N(1535)$S_{11}$ resonance there is little variation between the various theoretical results, for the ratio $A^p_{3/2}/A^p_{1/2}$ of the N(1520)$D_{13}$ resonance there is a large spread in values. Note also the large discrepancy between the photocouplings obtained from electro-photoproduction \cite{PDG} and the new values determined from eta-photoproduction \cite{Nimai1,Nimai2,Tiator}. Finally, in Fig.~\ref{n1535} we show the N(1535)$S_{11}$ proton helicity amplitude $A^p_{1/2}$ as a function of $Q^2$, for which there exist interesting new data \cite{Armstrong} (diamonds). The other points are obtained from a reanalysis of old(er) data, but now using the same values of the resonance parameters for all cases (from a compilation in \cite{Armstrong}). The solid curve represents the results of the collective algebraic model of \cite{BIL}, which were obtained by introducing a mixing angle of $\theta = -38$ degrees, but which do not contain the effects of meson-cloud couplings. We find good overall agreement with the data for the entire range of $Q^2$ values. \section{Summary and conclusions} We presented a simultaneous analysis of the four elastic form factors of the nucleon and the transition form factors in a collective model of baryons, and found in general good agreement with the data. The elastic electromagnetic form factors of the nucleon were studied for the space-like region $0 \leq Q^2 \leq 10$ (GeV/c)$^2$. Whereas for low $Q^2$ the form factors satisfy to a good approximation form factor scaling, in the region of high $Q^2$ they exhibit deviations from these simple scaling relations, most notably for the ratio of the neutron and proton magnetic form factors and the neutron and proton cross sections. The deviations of the nucleon form factors at low $Q^2$ from the dipole form were attributed to couplings to the meson cloud. For a phenomenological approach (as the present one) a good data set is a prerequisite. Whereas the proton form factors are relatively well known, there is still quite some controversy about the neutron form factors. New measurements of the polarization asymmetry in which the ratio of the electric and magnetic form factor of the neutron is extracted may help to further clarify the experimental situation. In conclusion, the present analysis of electromagnetic couplings shows that the collective model of baryons provides a good overall description of the available data. \section*{Acknowledgements} This work is supported in part by DGAPA-UNAM under project IN101997 and by grant No. 94-00059 from the United States-Israel Binational Science Foundation (BSF), Jerusalem, Israel.
\section{The proposed physical model} Let $A$ be the electromagnetic vector potential, and let the corresponding electromagnetic field be denoted by $ F_A = dA$. Further, let $f$ be a real valued function. Consider the following system of equations \begin{equation} dF_{A}=0 \label{syst0} \end{equation} \begin{equation} \delta (fF_{A})=0 \label{syst1} \end{equation} \begin{equation} -\bigtriangleup f +|F_{A}|^{2}f=\nu f. \label{syst2} \end{equation} The main goal of this paper is to describe a vortex-lattice type solution of this system of equations in two dimensions (cf. figure) by means of a finite difference approach. Results obtained in this way suggest that in the continuous domain limit such a solution consists of a continuous function $f$ and a connection $A$ whose curvature $2$-form $F_A$ is also continuous. Moreover, the equations are satisfied in the classical sense almost everywhere. Note that the displayed numerical solutions are not defined at the vortex points. (Here, the horizontal axes are indexed by discrete lattice points.) Moreover, their scaling should be matched to a given physical system. As explained in \cite{sow1} (and in a simplified Abelian version in \cite{sow2}), the particular form of this system is suggested by the geometry of principal $U(1)$-bundles. Purely phenomenologically (\ref{syst0})-(\ref{syst2}) can be motivated by evoking heuristics used frequently in the optical literature. Namely, it is often assumed in optics that the nonlinearities arising in the interaction of radiation with matter can be accounted for by representing $f$ as a series in tensorial powers of $F_A$ with coefficients characteristic of the material. Subsequently, one attempts to deduce properties of the coefficients from a microscopic theory. The shift of paradigm in this paper consists in assuming that $f$ depends in a geometrically invariant way on $F_A$ via equation (\ref{syst2}), and $f$ remains essentially independent of the material except for a simple scaling. We will see below that in contrast to the soft optical nonlinearities, (\ref{syst0})-(\ref{syst2}) cannot be understood as a small perturbation of a linear system. However, interaction of radiation with matter is described from the ``point of view'' of the former, and the electronic processes inside matter are never discussed directly. In order to explain superconductivity by means of a microscopic theory, one needs to display a mechanism that will let fermions overcome the obstacle imposed by the Pauli exclusion principle. A solution given by the famous BCS theory is based on the observation that when thermic noise is sufficiently low, it is energetically favored for electrons to join in pairs, known as Cooper pairs, which behave like bosons. The BCS theory is in a certain correspondence to the Ginzburg-Landau equations. These are nonlinear equations for a complex valued function $\psi$, often interpreted as an order parameter governing collective behavior of Cooper pairs. Periodic solutions of these equations in the form of vortices have been found by Abrikosov in \cite{abr}. It must be emphasized that strict validity of the BCS/GL approach, at least in its classical s-wave pairing version, is limited to low temperature superconductivity of metals. On the other hand, type II superconductivity is known to occur in materials structurally different from metals, like YBCO, and at relatively high critical temperatures. As many researchers pointed out, this suggests that mechanisms other than those encompassed by the BCS theory may be responsible for high temperature superconductivity. Those aspects of solid state theory that go beyond BCS seem particularly attractive in terms of the possibility of merging with the nonlinear Maxwell equations. It is possible that the new mathematical pattern introduced in this paper will be helpful in the description of the interaction of magnetic fields with composite particles. For illustration, consider the proposition that $f$ describes a locally varying {\em filling factor}. In this interpretation, a part of the field gets entrapped in composite bosons, composite fermions, and Laughlin quasiparticles, which in turn see only the remaining $\frac{1}{f}$-fraction of the field. If $f$ is a constant, this allows one to replace an electron picture with a composite particle picture, a suggestion that was present in science already ten years ago. However, if $f$ is a vortex-type solution, the composite particles will feel the vortex in $F_A$, which should induce Josephson-type effects. Thus, while microscopic theory is always constructed with an {\em a priori} fixed filling factor, (\ref{syst0})-(\ref{syst2}) would reflect the behavior of magnetic fields on a coarser scale. The above is meant to evoke some of the basic notions and ideas present in modern materials science. More thorough reviews can be found in the articles featured in the very incomplete list of references below (\cite{andr1}-\cite{zhang}). \section{Mathematics of a finite difference approximation} Consider the system (\ref{syst0})-(\ref{syst2}) on a two-dimensional flat torus $T^2$. In this case $\delta (fF)=0$ implies $\star F=\frac{B}{f}$ for a constant $B$. In addition, $dF=0$ and $F$ is the curvature of a certain connection $A$, provided its cohomology class satisfies $[F]\in 2\pi Z$. Thus, the system of equations reduces to \begin{equation} -\bigtriangleup f +\frac{B^2}{f}=\nu f , \label{tor1} \end{equation} and \begin{equation} \int_{T^2}\frac{B}{f}dV = 2\pi K \qquad\mbox{for an integer}\quad K. \label{tor2} \end{equation} Suppose $f$ is a solution of the first equation with parameter $B$. Then for any $c>0$, the function $cf$ is a solution of (\ref{tor1}) with $B$ replaced by $cB$. At the same time, this rescaling does not affect (\ref{tor2}) in any way, since the ratio $B/f$ remains fixed. However, the system behaves differently with respect to rescaling of the independent variable. Indeed, suppose $f$ satisfies both equations with parameters $B$, $\nu$, and $N$, then defining $g(x,y) = f(cx,cy)$, we have that $g$ satisfies (\ref{tor1}) and (\ref{tor2}) with parameters $Bc$, $\nu c^2$, and $N/c$, while its period is $1/c$ in both directions. This is consistent with the experimental fact that as a magnetic field normal to the surface increases, vortices should eventually collide with one another. (In physical reality they will at that point disappear together with the superconducting state). This fact is also important mathematically, as it allows us to first obtain a solution of (\ref{tor1}) with, say, $\int_{T^2}\frac{B}{f}dV < 2\pi$, and then rescale the independent variable to satisfy (\ref{tor2}). It appears that the system (\ref{tor1})-(\ref{tor2}) does not subdue itself to the standard techniques of variational calculus or topological analysis. In particular, perturbative methods do not apply to (\ref{tor1}), and no solutions arise as a result of bifurcation. In fact, in view of the theorem below and the results of numerical simulations, solutions are objects very unlike the familiar vortex solutions of nonlinear PDEs. To give some indication of the difficulties involved, consider the following. The equation (\ref{tor1}) is the Euler-Lagrange equation for the Lagrangian $L(f)= (\frac{1}{2}\int |\nabla f|^2 + B^2\int \ln(f))/(\int f^2)$. However, this functional is neither bounded below nor above. Indeed, let us note that $L(c_n)\rightarrow -\infty$ for constants $c_n \rightarrow 0$. On the other hand, let $f_n(x,y) = 1+\varepsilon + cos(2\pi n x)$. Since $\ln f_n(x,y) \geq \ln \varepsilon$, one easily checks that in this case $L(f_n)\rightarrow \infty$. Therefore, the best we can expect is to discover {\em local} extrema. Additional difficulty stems from the fact that since (\ref{tor1}) always admits a trivial constant solution, we must devise a method of telling trivial and nontrivial solutions apart. We will now consider a finite difference model of the system (\ref{tor1})-(\ref{tor2}). It is proven below that non-constant solutions of the discrete problem exist. The proof is independent of the number of points in the discretization ($n^2$) but relies on finite-dimensionality essentially, and does not admit a direct generalization to the analytic case. However, all the universal parameters used in the proof, like the $L_2$-norm of $f$ and $B$, are asymptoticly independent of $n$. Thus, we conjecture existence of the continuous domain solutions of (\ref{tor1}) that satisfy the equation a.e. in the classical sense and retain the particular vortex morphology. It is convenient to introduce the following notation. $\int = \frac{1}{n^2}\sum\limits_i\sum\limits_j$, $\int\limits_ o = \frac{1}{n^2}\sum\limits_{i\neq i_0}\sum\limits_{j\neq j_0}$, where indices $i,j$ run through the discrete $n$-by-$n$ lattice. Also, $\bigtriangleup$ denotes the common five-point periodic discrete Laplacian, i.e. $ \bigtriangleup f(\frac{i}{n},\frac{j}{n}) = n^2 \left(f(\frac{i+1}{n},\frac{j}{n})+f(\frac{i}{n},\frac{j+1}{n}) +f(\frac{i-1}{n},\frac{j}{n})+f(\frac{i}{n},\frac{j-1}{n}) -4f(\frac{i}{n},\frac{j}{n})\right) $ and $\nabla = (\frac{\partial}{\partial x},\frac{\partial}{\partial y})$ is the simplest two-point periodic gradient, i.e. say $ \frac{\partial}{\partial x}f(\frac{i}{n},\frac{j}{n}) = n (f(\frac{i+1}{n},\frac{j}{n})-f(\frac{i}{n},\frac{j}{n})). $ In particular, the discrete integration-by-parts formula holds, i.e. $-\int(\bigtriangleup f)g = \int (\nabla f, \nabla g)$. Consider the function \[ \Phi(f) = \frac{1}{2}\int |\nabla f|^2 + B^2\int \ln(f). \] Pick arbitrary real numbers $a,b,c$, fix a point $(x_0,y_0)=(\frac{i_0}{n},\frac{j_0}{n})$, and a number \begin{equation} m_n = \left(\frac{n^2}{b}-\frac{1}{c}\right)^{-1}. \label{the_m} \end{equation} Two submanifolds in $R^{n^2}$, \begin{equation} \label{D} D^n_{a,b,c} = \{f>0: \int f^2 = a^2, \int\frac{1}{f} \leq \frac{1}{b}, \min f = f(x_0,y_0) = m_n \} \end{equation} and its boundary \begin{equation} \label{delD} \partial D^n_{a,b,c} = \{f>0: \int f^2 = a^2, \int\frac{1}{f} = \frac{1}{b}, \min f = f(x_0,y_0) = m_n \}, \end{equation} play a fundamental role in understanding the nature of critical points of $\Phi$. Depending on the particular value of $a$, $b$, and $c$, the set $D^n_{a,b,c}$ is either empty, an $(n^2-2)$-dimensional spherical disk, or it degenerates to a point. Consider the hyper-plane $H_n = \{f:f(x_0,y_0) = m_n \}$. $D^n_{a,b,c}$ is a spherical disk immersed in $H_n$ precisely when $V$, the point closest to the origin of an open submanifold given by $\int\limits_o\frac{1}{f} = \frac{1}{b} - \frac{1}{n^2m_n}$, is located inside the ball $\int\limits_o f^2 \leq a^2-\frac{m_n^2}{n^2}$. One easily finds $V(x,y) = const = (n^2-1)c$ for all $x\neq x_0, y\neq y_0$, and for it to be inside the ball, it is necessary and sufficient that $a$ and $c$ satisfy \begin{equation} \frac{(n^2-1)^3}{n^2}c^2 < a^2 - \frac{m_n^2}{n^2}. \label{a_and_c} \end{equation} Conversely, if condition (\ref{a_and_c}) holds then $D^n_{a,b,c}$ is a nonempty spherical disk. Formally, one needs to check in addition that $m_n$ is indeed a minimum of every function that satisfies the two other conditions in (\ref{D}), but this is straightforward. At this point I would like to point out that in the description of local minima of $\Phi$ below, the parameter $a$ is physical, and with good faith can be regarded as the $L_2$-norm of the critical point, whereas the parameters $b,c$ are auxiliary and will converge to $0$ as the density of discretization $n$ increases to infinity. In particular, interpreting $\frac{1}{b}$ as an approximate value of the integral of the reciprocal of the function where $\Phi$ attains its local minimum is erroneous since the function develops a singularity at $(x_0,y_0)$. The following theorem will be proven. \begin{th} Fix constants $B$ and $a$ as above. For a certain choice of the constants $b=b(n,a)$, and $c=c(n,a)$, the function $f\rightarrow\Phi(f)$ assumes local relative minima in $D^n_{a,b,c}$. In particular, the corresponding critical point, say $f_0>0$, satisfies the finite difference version of (\ref{tor1}) everywhere except one point, i.e. \begin{equation} -\bigtriangleup f_0(x,y) +\frac{B^2}{f_0(x,y)}=\nu f_0(x,y).\qquad \mbox{for all}\quad (x,y) \neq (x_0,y_0). \label{tor_dscr} \end{equation} Moreover, if $B$ is sufficiently large, then $f_0$ is not a constant function. \label{minima} \end{th} {\em Proof.} It is convenient to first treat $D^n_{a,b,c}$ formally, and check that condition (\ref{a_and_c}) is satisfied later. Proceeding formally, let $N$ denote an outward normal vector to $\partial D^n_{a,b,c}$ inside its ambient sphere, i.e. $N$ is tangent to the $(n^2-2)$-dimensional sphere $S_a = \{f:\int f^2 = a^2,\quad f(x_0,y_0) = m_n\}$ and points away from the region $D^n_{a,b,c}$. The main task is to show that $N\Phi>0$. It will then follow from smoothness of $\Phi$ ($\nabla$ is a linear operator and $f\rightarrow ln(f)$ is smooth for $f>0$) that it assumes a local minimum inside $D^n_{a,b,c}$. Let us choose for $N$ the vector field defined by \begin{equation} N_f(x,y) = \left\{\begin{array}{ll} \frac{1}{ba^2}f(x,y) -\frac{1}{f(x,y)^2} &\qquad \mbox{for}\quad (x,y) \neq (x_0,y_0)\\ 0 &\qquad \mbox{otherwise}, \end{array}\right . \end{equation} where $f\in\partial D^n_{a,b,c}$. A direct calculation shows that \[ N_f\Phi(f) = \frac{1}{ba^2} \left(-\int\limits_ of\bigtriangleup f + B^2(1-\frac{1}{n^2})\right) + \int\limits_ o\frac{1}{f^2}\bigtriangleup f - B^2\int\limits_ o\frac{1}{f^3}. \] It remains to analyze terms one by one. First, since a function in $D^n_{a,b,c}$ assumes its minimum at $(x_0, y_0)$, we obtain \begin{equation} -\int\limits_ of\bigtriangleup f = \int|\nabla f|^2 +\frac{1}{n^2}\bigtriangleup f(x_0,y_0)m_n \geq \int|\nabla f|^2 \geq 0. \label{first} \end{equation} Next, by definition of $D^n_{a,b,c}$ and (\ref{the_m}) we obtain \[ \frac{1}{b} = \int\frac{1}{f} = \int\limits_ o\frac{1}{f}+ \frac{1}{n^2}\frac{1}{m_n} = \int\limits_ o\frac{1}{f}+ \frac{1}{b} - \frac{1}{n^2c}, \] so that $\sum\limits_{i\neq i_0}\sum\limits_{j\neq j_0}\frac{1}{f} = \frac{1}{c}$ and therefore \begin{equation} f(x,y) > c \qquad \mbox{for}\quad (x,y) \neq (x_0,y_0). \label{lower_bound} \end{equation} As an immediate application we obtain \begin{equation} \int\limits_ o\frac{1}{f^3} \leq \frac{1}{c^3}(1-\frac{1}{n^2}). \label{second} \end{equation} Finally, we obtain the following inequality \begin{equation} \begin{array}{llll} && \int\limits_ o\frac{1}{f^2}\bigtriangleup f\\ &=& \frac{1}{n^2}\sum\limits_{i\neq i_0}\sum\limits_{j\neq j_0} \frac{n^2}{f(\frac{i}{n},\frac{j}{n})^2}\left(f(\frac{i+1}{n},\frac{j}{n})+f(\frac{i}{n},\frac{j+1}{n}) +f(\frac{i-1}{n},\frac{j}{n})+f(\frac{i}{n},\frac{j-1}{n}) -4f(\frac{i}{n},\frac{j}{n})\right) \\ &\geq& \sum\limits_{i\neq i_0}\sum\limits_{j\neq j_0}\frac{1}{f(\frac{i}{n},\frac{j}{n})^2} (-4f(\frac{i}{n},\frac{j}{n})) = -4\sum\limits_{i\neq i_0}\sum\limits_{j\neq j_0}\frac{1}{f(\frac{i}{n},\frac{j}{n})} \\ &=& -4n^2(\frac{1}{b}-\frac{1}{n^2}\frac{1}{m_n}) = -\frac{4}{c}. \end{array} \label{third} \end{equation} Together, estimates (\ref{first}), (\ref{second}), (\ref{third}) yield \begin{equation} N_f\Phi(f) \geq (1-\frac{1}{n^2})B^2(\frac{1}{ba^2}-\frac{1}{c^3}) - \frac{4}{c}. \label{Nf_est} \end{equation} So far no assumption has been made about the constants. Now, in order to guarantee existence of local minima one needs to ensure that $D^n_{a,b,c}$ is nonempty, and that the outer derivative is positive. Both these requirements are met if we pick \[ c = c(n) = \frac{a}{2n^2}, \] and \[ b = b(n) = \frac{a}{16n^6}. \] With this choice of $c$, (\ref{a_and_c}) holds and if $f$ satisfying the first two conditions in (\ref{D}), assumed at some other point a value at least as small as $m_n$, we would have $\int\limits_o\frac{1}{f} > \frac{2m_n^{-1}}{n^2} = \frac{2}{b}-\frac{1}{cn^2} > \frac{1}{b}$, which is a contradiction. Thus $D^n_{a,b,c}$ is nonempty. On the other hand, inequality (\ref{Nf_est}) becomes $ N_f\Phi(f) \geq \frac{8B^2}{a^3}(1-\frac{1}{n^2})n^6 - \frac{8}{a}n^2$ which implies $ N_f\Phi(f)>0$ for $n$ sufficiently large. Consequently, $\Phi$ assumes a local minimum inside $D^n_{a,b,c}$. Equation (\ref{tor_dscr}) is automatically satisfied because it expresses the fact that the component of the derivative of $\Phi$ which is tangent to the sphere $\int f^2 = a^2$ vanishes in all directions except possibly the $(x_0,y_0)$-direction. Still, the local minimum $f_0$ could {\em a priori} be a function constant everywhere, except at the discontinuity in $(x_0,y_0)$. This can be avoided by taking $B$ sufficiently large. As $B$ increases, the solution which is constant except at $(x_0,y_0)$, say \[ f_1(x,y)=\left\{\begin{array}{r@{\quad:\quad}l} \kappa & (x,y)\neq(x_0,y_0) \\ m_n & \mbox{otherwise} \end{array}\right. \] becomes unstable, i.e. it corresponds to a saddle point on the graph of $\Phi$. Indeed, one checks directly that \[ \frac{d^2}{d\varepsilon ^2}\Phi(f_1+\varepsilon\phi ) = \int|\nabla \phi|^2 - B^2\int \frac{\phi ^2}{f_1^2}, \] for $\phi$ tangent to the origin-centered sphere at $f_1$, so that $\int f_1\phi = 0$. Now consider $\phi$ to be a non-constant eigenfunction of the discrete Laplacian on a torus and let $\lambda$ denote the corresponding eigenvalue. Shifting it if necessary, we can assume that $\phi(x_0,y_0)=0$, so that $\phi$ is orthogonal to $f_1$. We have that $\int|\nabla \phi|^2 = \lambda\int\phi ^2$, and thus $\frac{d^2}{d\varepsilon ^2}\Phi(f_1+\varepsilon\phi ) < 0$ only if $\frac{B^2}{\kappa ^2} > \lambda$. Thus, $f_1$ is not a local minimum for $B$ sufficiently large. Moreover, since $\int f_1^2 = a^2$, $\kappa$ converges to $a$ as $n$ increases. Naturally, one can make sure that $\lambda$ remains fixed regardless of $n$ by always picking $\phi$ to be a discretization of the same trigonometric function. This shows that a choice of $B$ which guarantees that $f_0$ is a nontrivial solution does not depend on the discretization. $\Box$ \section{Closing remarks} The proof above depends essentially on the fact that all functions and manifolds are discrete. Indeed, the constants $b=b(n), c=c(n)$ we have used tend to infinity as $n\rightarrow\infty$, and some of the estimates make no sense in the limit. However, it is important that the magnetic induction $B$ and the $L_2$-norm of $f$ do not depend on $n$. On the other hand, simulation shows that the discrete solution $f_0$ retains its particular morphology independently of $n$, and is always subharmonic. In addition, multiplying (\ref{tor_dscr}) by $f\chi_{\{(x,y) \neq (x_0,y_0)\}}$ and inspecting the vicinity of the singularity we easily obtain $\nu \sim \frac{1}{a^2}\left(B^2 +\int |\nabla f|^2\right)$, which is also expected to converge. In summary, we have enough evidence to believe that the figure above shows a good approximation to a strong solution of the continuous version of equation (\ref{tor1}) that would posess a vortex of Lipschitz regularity. It is also consistent with the one-dimensional case, where solutions of the analogous nonlinear equation can be expressed in terms of a closed-form integral. I point out for completeness that continuity of such a positive solution $f_0$ guarantees continuity of $F_A = \star\frac{B}{f_0}$, and by rescaling the independent variables we can satisfy the condition that the cycle of $F_A$ be an integral multiple of $2\pi$. This is sufficient to solve $dA=F_A$ for $A$ on a compact surface, and the solution $A$ has sufficient regularity to retain its geometric interpretation as a connection $1$-form. On the other hand, $A$ contains all the information necessary to derive the basic tenets of superconductive electronics, like the Josephson effect. It is important to realize that none of the propositions stated above necessarily rely on the lattice being a simple square lattice. Most likely, a hexagonal lattice setting would yield the same qualitative results. The only reason for using a square lattice is to avoid overwhelming numerical complexity in experiments, as well as arithmetical nuisance in theory. I introduced the system (\ref{syst0})-(\ref{syst2}) in 1993, and later investigated its properties in my thesis. Some of those early results are contained in \cite{sow1}. The geometry was first given a loose but essentially correct physical interpretation in \cite{sow2}. This work is a departure from the bulk of current research I conduct in collaboration with Professor R. R. Coifman in the area of Computational Harmonic Analysis. Working with Raphy is a stimulating adventure, and I want to thank him here for enhancing my understanding of mathematics. I also wish to thank my friend Fred Warner for proofreading the manuscript.
\section{Introduction} Supersymmetry (SUSY) breaking is very important to connect superstring theory to low-energy physics. Induced soft SUSY breaking terms are significant to currently running and future experiment, and they also have cosmological implications. Actually, soft SUSY breaking terms have been derived, e.g. within the framework of weakly coupled heterotic string theory under the assumption that only $F$-terms of the dilaton $S$ and moduli fields $T^i$ contribute to SUSY breaking \cite{CCM,ST-soft,BIM,multiT}. Also various phenomenological aspects of these string-inspired soft terms have been studied. Recently strongly coupled superstring theories as well as strongly coupled SUSY gauge theories have been studied. Strongly coupled $E_8 \times E_8$ heterotic string theory has been considered as M-theory compactified on $S^1/Z_2$ \cite{M-theory}. In particular, the 4-dimensional effective theory has been obtained and supersymmetry (SUSY) breaking has been studied \cite{banks}-\cite{ELP}. M-inspired soft SUSY breaking parameters have a certain difference from those derived from weakly coupled string models. Soft parameter formulae between the weakly and strongly coupled regions are connected by the parameter $\alpha(T+\bar T)$. Actually, several phenomenological aspects of soft SUSY breaking parameters have been investigated, e.g. radiative electroweak symmetry breaking and s-spectrum \cite{BKL}, although the large $\tan \beta$ region has not been studied extensively. The standard embedding leads to the gauge group $E_6 \times E_8'$, where the top, bottom and tau Yukawa couplings are unified, that is, we have a large value of $\tan \beta$. In this sense, the large $\tan \beta$ scenario is one natural target to study. When $E_6$ breaks into the standard model (SM) gauge group $G_{SM}=SU(3)_C\times SU(2)_L \times U(1)_Y$, another type of soft scalar terms appear, i.e., $D$-term contributions. On the other hand, non-standard embedding including Wilson line breaking is also possible \cite{nonst} and that can lead to smaller gauge groups, e.g. $SU(3)\times SU(2) \times U(1)$ or $SU(5)$. Thus, in non-standard embedding a small value of $\tan \beta$ can also be realized. Furthermore, non-standard embedding, in general, has 5-branes, that is, in 4-dimensional effective field theory we have a new type of moduli fields $Z^n$ whose vacuum expectation values (VEVs) correspond to positions of 5-branes $z^n$ within $[0,1]$ along the orbifold $S^1/Z_2$. It is possible that $F$-terms of the 5-brane moduli fields $Z^n$ also contribute to SUSY breaking. Recently 4-dimensional effective theory with $Z^n$ has been studied \cite{5brane} and SUSY breaking terms have also investigated \cite{5b-soft}. The purpose of this paper is to take into account several effects which are mentioned above and to investigate successful electroweak breaking conditions and the lightest superparticle (LSP) systematically. In the whole paper, we assume that the massless particle content below the grand unified theory (GUT) scale is exactly the same as the minimal supersymmetric standard model (MSSM). In the large $\tan \beta$ scenario, the stau (mass)$^2$ becomes small and negative because of large and negative radiative corrections from the large Yukawa coupling. Thus the stau mass leads to a significant constraint for large $\tan \beta$. We cover the whole realistic region of $\tan \beta$ including large $\tan \beta$. We investigate $\alpha(T+\bar T)$-dependence of electroweak symmetry breaking and s-spectrum. Furthermore, we study the effect due to $F$-terms of the moduli fields $Z^n$. This paper is organized as follows. In section 2, after we describe the structure of the soft parameters in M-theory without 5-brane, we investigate the electroweak symmetry breaking conditions and the positivity of stau mass squared under the assumption of top-bottom-tau Yukawa coupling unification, that is, the large $\tan \beta$ scenario. $D$-term contributions, which appear in $E_6$ breaking into $G_{SM}$, are also taken into account. We also consider the case with small $\tan \beta$. In section 3, we consider soft parameters in M-theory that the 5-brane moduli field also contribute to SUSY breaking. In this case, electroweak symmetry breaking and mass spectra are studied and allowed regions are compared with the case without 5-brane effects. Section 4 is devoted to conclusions. \section{SUSY breaking in M-theory without 5-branes} In this section we study SUSY breaking in M-theory without 5-branes. For example, standard embedding corresponds to this class. We discuss the soft breaking parameters derived within this framework and study their phenomenological implications. \subsection{Soft parameters} The 4-dimensional effective supergravity action of M-theory can be analyzed by expanding it in powers of the two dimensionless variables: $\varepsilon_1=\kappa^\frac{2}{3}\pi\rho/V^\frac{2}{3}$ and $\varepsilon_2=\kappa^\frac{2}{3}/\pi\rho V^\frac{1}{3}$, likewise the weakly coupled heterotic string theory expanded by the string coupling $\varepsilon_s=e^{2\phi}/(2\pi)^5$ and the worldsheet sigma model coupling $\varepsilon_\sigma=4\pi\alpha'/V^{1/3}$, where $\kappa^2$, $\rho$ and $V$ denote the 11-dimensional gravitational coupling, the 11-dimensional length and the Calabi-Yau volume respectively. Here we consider only the overall moduli field $T$. Thus, the 4-dimensional effective supergravity includes the dilaton $S$ and the moduli field $T$ in addition to chiral multiplets $\Phi^p$, gauge multiplets and the graviton multiplet. Scalar components of $S$ and $T$ can be identified as \begin{eqnarray} Re(S)&=&\frac{1}{2\pi}(4\pi \kappa^2)^{-2/3}V \nonumber ,\\ Re(T)&=&\frac{6^{1/3}}{8\pi}(4\pi \kappa^2)^{-1/3}\pi\rho V^{1/3}. \end{eqnarray} The K\"{a}hlar potential is computed by M-theory expansion for matter fields up to order of $\kappa^{4/3}$ \cite{LOW}, \begin{eqnarray} K&=&-\log(S+\bar{S})-3\log(T+\bar{T})+\left(\frac{3}{T+\bar{T}} +\frac{\alpha}{S+\bar{S}}\right)|\Phi|^2 \label{K}. \end{eqnarray} Similarly the gauge kinetic functions of the observable and hidden sectors are calculated \cite{banks,choi1,nilles1,nilles2}, \begin{eqnarray} f_{E_6}&=&S+\alpha T, \hspace{6pt} f_{E_8}=S-\alpha T \label{f6}, \end{eqnarray} where $\alpha={1}/{(8\pi^2)}\int \omega \wedge \left[tr(F \wedge F)-\frac{1}{2}tr(R \wedge R)\right]$. In addition, the tree-level superpotential is obtained, \begin{eqnarray} W&=&Y_{pqr}\Phi^p\Phi^q\Phi^r \label{W}. \end{eqnarray} Here $\alpha$ is of the order one at the ``physical'' point where gauge coupling unification can be realized \cite{banks}. Note that the superpotential $W$ and gauge kinetic functions $f_{E_6}$ and $f_{E_8}$ are exact up to nonperturbative corrections, while there are small additional perturbative corrections to the K\"{a}hlar potential which are of order $1/4\pi^2Re(S)$ or $1/[4\pi^2Re(T)]^3$ $\sim O(\alpha_{GUT}/\pi)$ in the strong coupling limit $\varepsilon_s \gg 1$. Now we know the forms of the K\"ahler potential $K$, the gauge kinetic function $f_{E_6}$ and the tree-level superpotential $W$. Thus, we can derive the formula of soft SUSY breaking parameters following ref.\cite{SW,ST-soft}. We assume that a nonperturvative superpotential of $S$ and $T$ is induced and $F$-terms of $S$ and $T$ contribute to SUSY breaking. The gravitino mass $m_{3/2}$ is obtained \begin{equation} 3m_{3/2}^2 = \frac{|F^S|^2}{3(S+\bar{S})^2} +\frac{|F^T|^2}{(T+\bar{T})^2}-V_0, \end{equation} where $V_0$ is the vacuum energy. Hence, following ref.\cite{BIM}, we parameterize $F$-terms, \begin{eqnarray} F^S&=&\sqrt{3}m_{3/2}C(S+\bar{S})\sin\theta e^{-i\gamma_S} \label{Fs},\\ F^T&=&m_{3/2}C(T+\bar{T})\cos\theta e^{-i\gamma_T},\label{Ft} \end{eqnarray} where $C^2=1+V_0/3m_{3/2}$. Using these parameters, we can write the soft parameters, i.e., the gaugino mass $M_{1/2}$, the soft scalar mass $m$ and the $A$-parameter as follows \cite{CKM} \begin{eqnarray} M_{1/2}&=&\frac{\sqrt{3}Cm_{3/2}}{(S+\bar{S})+\alpha(T+\bar{T})} \left((S+\bar{S}) \sin\theta e^{-i\gamma_S} \right. \nonumber \\ &+& \left. \frac{\alpha(T+\bar{T})}{\sqrt{3}}\cos\theta e^{-i\gamma_T}\right) \label{Mg},\\ m^2&=&V_0+m^2_{3/2}-\frac{3C^2m^2_{3/2}}{3(S+\bar{S})+\alpha(T+\bar{T})} \nonumber\\ &{}& \times \left\{ \alpha(T+\bar{T})\left(2-\frac{\alpha(T+\bar{T})}{3(S+\bar{S}) +\alpha(T+\bar{T})}\right)\sin^2\theta \right. \nonumber\\ &{}& \left. +(S+\bar{S})\left(2-\frac{3(S+\bar{S})}{3(S+\bar{S}) +\alpha(T+\bar{T})}\right)\cos^2\theta \right.\nonumber\\ &{}& \left. -\frac{2\sqrt{3}\alpha(T+\bar{T})(S+\bar{S})}{3(S+\bar{S}) +\alpha(T+\bar{T})}\sin\theta\cos\theta\cos(\gamma_S-\gamma_T) \right\} \label{ms},\\ A&=&\sqrt{3}Cm_{3/2} \left\{ \left(-1+\frac{3\alpha(T+\bar{T})}{3(S+\bar{S}) +\alpha(T+\bar{T})}\right)\sin\theta e^{-i\gamma_S} \right.\nonumber\\ &{}& \left. +\sqrt{3}\left(-1+\frac{3(S+\bar{S})}{3(S+\bar{S}) +\alpha(T+\bar{T})}\right)\cos\theta e^{-i\gamma_T} \right\}. \label{A} \end{eqnarray} In following analyses we concentrate to the case with the vanishing vacuum energy, $V_0=0$ ($C=1$) and no CP phase $\gamma_S=\gamma_T=0$. Since the gauge kinetic function $f_{E_6}$ provides the gauge coupling constant, that is, $Ref_{E_6}=1/g^2_{E_6}$, the phenomenological requirement $g^{-2}_{GUT} \simeq 2$ leads to \begin{eqnarray} (S+\bar{S})+\alpha(T+\bar{T}) \simeq 4 \label{cons}. \end{eqnarray} Moreover, the constraint for $Ref_{E_8} > 0$ leads to $(S+\bar{S}) >\alpha(T+\bar{T})$. Therefore, we have \begin{eqnarray} 0 < \alpha(T+\bar{T}) \mathrel{\mathpalette\gs@align<} 2 \label{lam}. \end{eqnarray} Note that there are three free parameters $m_{3/2}$, $\theta$ and $\alpha(T+\bar{T})$, and in particular, $\alpha(T+\bar{T})$ is the characteristic parameter of M-theory. The limit $\alpha(T+\bar{T}) \to 0$ corresponds to weakly coupled theory. \subsection{Electroweak symmetry breaking and mass spectrum} We now analyze phenomenology of M-theory by using eqs.(\ref{Mg}),(\ref{ms}),(\ref{A}) as boundary conditions of renormalization group equations (RGEs) at the GUT scale. We investigate successful radiative electroweak symmetry breaking and positive masses squared for any sfermion. That gives constraints for the parameters $m_{3/2}$, $\alpha(T+\bar{T})$ and $\theta$. If we assume a certain type of $\mu$-term generation mechanism, we could fix magnitudes of the supersymmetric Higgs mixing mass $\mu$ and the corresponding SUSY breaking parameter $B$. However, we do not take such a procedure here. Because we would like to study generic case. We determine these magnitudes by using the following minimization conditions of the MSSM Higgs potential, \begin{eqnarray} m_1^2+m_2^2&=&-\frac{2 \mu B}{\sin\beta}, \nonumber \\ m_1^2-m_2^2&=&-(M_Z^2+m_1^2+m_2^2)\cos2\beta ,\label{min} \end{eqnarray} where $m_{1,2}^2=m_{H_1,H_2}^2+\mu^2$, and $m_{H_1,H_2}^2$ are soft scalar masses of Higgs fields. The Higgs fields for the down sector and the up sector are denoted by $H_1$ and $H_2$. Using these equations, $\mu$ and $B$ are written in terms of the soft scalar masses and $\tan \beta$. Realization of successful electroweak symmetry breaking leads to the condition, \begin{eqnarray} m_1^2 m_2^2 < |\mu B|^2, \label{EW} \end{eqnarray} and the bounded-from-below condition along the $D$-flat directions in the Higgs potential, that is, \begin{eqnarray} m_1^2+m_2^2 > 2|\mu B|, \end{eqnarray} should be satisfied in order for stability of the potential along the $D$-flat direction. In the large $\tan \beta$ scenario, these conditions lead to \begin{eqnarray} m_{H_1}^2-m_{H_2}^2 \mathrel{\mathpalette\gs@align>} M_Z^2/2 .\label{EW2} \end{eqnarray} Also, we require that a physical mass squared should be positive for any sfermion. Since the sleptons have no large positive-definite radiative corrections from loops involving the gluino, the sleptons become lighter than squarks at low energy. In particular, in the case of large tan$\beta$, the stau mass squared has a sizable and negative radiative correction due to the large Yukawa coupling. lighter stau mass becomes significant for large $\tan \beta$ \footnote{ For example, in refs.\cite{kkmz,kks} the constraint $m_{\tilde{\tau}_1}^2>0$ has been considered in details for (finite) $SU(5)$ GUTs and $SO(10)$ GUTs.}. Furthermore, what is the LSP is an important issue. One candidate is the lightest neutralino and another candidate is the lighter stau for large $\tan \beta$ from the above reason \footnote{In ref.\cite{stau} cosmological implications of the stau LSP have been discussed. The stau LSP could not be a candidate for cold dark matter. In the case with the stau LSP we need another candidate for cold dark matter.}. The stau mass matrix is \begin{eqnarray} {\bf m_{\tilde{\tau}}^2} = \left( \begin{array}{cc} m_{\tilde{\tau}_L}^2+(-\frac{1}{2}+\sin^2\theta_W)M_Z^2 \cos2\beta & vY_\tau(A_\tau \cos\beta-\mu \sin\beta) \\ vY_\tau(A_\tau \cos\beta-\mu \sin\beta) & m_{\tilde{\tau}_R}^2-M_Z^2 \sin^2\theta_W \cos2\beta \end{array} \right) \label{taumtrx}, \end{eqnarray} where $v^2 \equiv <H_1>^2+<H_2>^2$, and we have neglected the tau (mass)$^2$ $M_\tau^2$ term here. The lighter stau mass squared is obtained as \begin{eqnarray} m_{\tilde{\tau}_1}^2&=&\frac{1}{2} \Biggl( m_{\tilde{\tau}_L}^2+m_{\tilde{\tau}_R}^2 -\frac{1}{2}M_Z^2 \cos2\beta \nonumber \\ &-& [(m_{\tilde{\tau}_L}^2-m_{\tilde{\tau}_R}^2 -(\frac{1}{2}-2\sin^2 \theta_W) M_Z^2 \cos2\beta )^2 \nonumber \\ &+& 4vY_\tau(A_\tau \cos\beta-\mu \sin\beta)^2]^{1/2} \Biggr). \label{tau1} \end{eqnarray} We take the following input parameters, \begin{eqnarray} M_\tau&=&1.777GeV,\hspace{12pt} M_Z=91.188GeV \nonumber, \\ \alpha_{EM}^{-1}(M_Z)&=&127.9+\frac{8}{9\pi}\log\frac{M_t}{M_Z} \nonumber, \\ \sin^2\theta_W&=&0.2319-3.03 \times 10^{-5}\Delta_t-8.4 \times 10^{-8}\Delta_t^2 , \label{input} \end{eqnarray} where $\Delta_t=M_t[GeV]-165.0$. Here $M_\tau$ and $M_t$ are physical tau lepton and top quark masses. We define the SUSY scale as the arithmetic average of the stop squared mass eigenvalues \cite{MSUSY}, \begin{eqnarray} M_{SUSY}^2=\frac{m_{\tilde{t_1}}^2+m_{\tilde{t_2}}^2}{2} \label{SUSY}. \end{eqnarray} \subsubsection{Standard embedding} First of all, we consider the case of standard embedding, where the gauge group $E_6 \times E_8'$ is obtained and the top, bottom and tau Yukawa couplings are unified at the GUT scale. From the experimental value of $M_\tau$, we obtain tan$\beta=53$ and $M_t=175$GeV. We assume at the GUT scale $E_6$ breaks into $G_{SM}$ and exactly the MSSM matter content is obtained. On top of that, we assume this symmetry breaking induces no extra contributions to soft parameters. Then we use eqs.(\ref{Mg}),(\ref{ms}),(\ref{A}) as the initial conditions of RGEs at the GUT scale. The latter assumption will be relaxed later. Before studying the constraints, it is useful to show how the soft parameters depend on the parameters, $2\alpha Re(T)$ and $\theta$. The $A$-parameter is always negative. Fig. 1 shows the ratios $-A/M_{1/2}$ and $m/M_{1/2}$ against $2\alpha Re(T)$ for values of $\theta$. The three solid curves show $-A/M_{1/2}$. The upper, intermediate and lower curves correspond to $-A/M_{1/2}$ for $\theta =0, \pi/5$ and $\pi/2$, respectively. Similarly, the two dotted curves show $m/M_{1/2}$. The upper and lower correspond to $m/M_{1/2}$ for $\theta =\pi/5$ and $\pi/2$, respectively. The weak coupling limit leads to \begin{equation} {A \over M_{1/2}} =-1, \quad {m \over M_{1/2}} = {1 \over \sqrt 3}, \label{sumrule} \end{equation} for any value of $\theta$ \footnote{These relations have an implication for finiteness and RG invariance \cite{JJ,kkmz}.}. We have $-A/M_{1/2} \approx -1$ for any value of $2\alpha Re(T)$ except the dilaton-dominant case $|\sin \theta| \approx 1$ and the moduli-dominant case $|\cos \theta| \approx 1$. In the dilaton-dominant (moduli-dominant) case, $-A/M_{1/2}$ becomes smaller (larger) in particular at $2\alpha Re(T) > 1.5$. The ratio $m/M_{1/2}=1/\sqrt 3$ holds approximately for any value of $2\alpha Re(T)$ except in the case with $|\sin \theta | > 0.9$ and the purely moduli-dominant case with $|\cot \theta| > O(10)$. In the moduli dominant case with $|\sin \theta | > 0.9$, as $2\alpha Re(T)$ increases, $m/M_{1/2}$ becomes small in particular at $2\alpha Re(T)> 1.5$. Thus, we have similar results in the region where the relations (\ref{sumrule}) hold approximately. Figs. 2 show the regions excluded by the electroweak symmetry breaking conditions and the constraint $m^2_{\tilde{\tau}_1} > 0$. The dotted region denotes the forbidden parameter region by the electroweak breaking conditions, while the squares correspond to $m^2_{\tilde{\tau}_1} < 0$. In addition the asterisks denote the region with $m^2_{\tilde{\tau}_1} < m^2_{\tilde{\chi}^0_1}$, where the LSP is the stau. The constraint due to successful electroweak symmetry breaking is relaxed more as $2\alpha Re(T)$ increases. In particular, a large $\theta$ within $[0,\pi]$ is favorable for successful electroweak symmetry breaking. On the other hand, in the case with a large $\theta$, we have $m^2_{\tilde{\tau}_1} < 0$ for large $2\alpha Re(T)$. That is because we have $m^2 < 0$ at the GUT scale by eq.(\ref{ms}), e.g. in the cases with $\theta=\pi/2$ and $3\pi/4$ for $\alpha(T+\bar{T}) > 1.8$ and $1.3$, respectively. In any allowed region, the LSP is the stau. \subsubsection{$D$-term contributions} We have assumed that symmetry breaking has no effect on soft parameters. However, if a gauge group symmetry breaks reducing its rank, additional contributions to soft scalar mass terms, in general, appear, i.e. $D$-term contribution \cite{Dterm1,Dterm2}. Thus, when the gauge group $E_6$ breaks to $G_{SM}$, the additional contributions to soft scalar mass are able to appear at the breaking scale. Here we take into such $D$-term contributions effects to soft scalar masses. In general, these $D$-term contributions are proportional to quantum numbers of broken diagonal generators. In the present case, we have two independent universal factors of $D$-term contributions, which we denote as $m_{D1}^2$ and $m_{D2}^2$, because the rank of $E_6$ is larger than $G_{SM}$ by two. We regard these as free parameters and then analyze their effects on the parameter space. The soft scalar masses at the GUT scale including $D$-term contributions are written as \begin{eqnarray} m_{\tilde{Q}}^2=m_{\tilde{t}}^2=m_{\tilde{\tau}_R}^2&=& m_{27}^2-m_{D1}^2+m_{D2}^2 , \nonumber\\ m_{\tilde{b}}^2=m_{\tilde{\tau}_L}^2&=& m_{27}^2+3m_{D1}^2+m_{D2}^2 , \nonumber\\ m_{H_1}^2&=& m_{27}^2-2m_{D1}^2-2m_{D2}^2 , \nonumber\\ m_{H_2}^2&=& m_{27}^2+2m_{D1}^2-2m_{D2}^2 ,\label{Dterm} \end{eqnarray} where $m_{27}^2$ is the soft scalar mass in eq.(\ref{ms}). We now analyze $D$-term contribution effects on the parameter space fixing $\tan \beta =53$, which has been obtained under the assumption of the top-bottom-tau Yukawa coupling unification. The term $m_{D1}^2$ with a negative and sizable value is helpful to create a positive gap $m_{H1}^2-m_{H2}^2$, which is favorable to successful electroweak symmetry breaking. Also, a negative value of $m_{D1}^2$ increases $m_{\tilde{\tau}_R}^2$, but it reduces $m_{\tilde{\tau}_L}^2$. The (mass)$^2$ $m^2_{\tilde{\tau}_R}$ is dominant to $m^2_{\tilde{\tau}_1}$ around the universal case with $m^2_{\tilde{\tau}_R}=m^2_{\tilde{\tau}_L}$. Therefore negative $m_{D1}^2$ may increase $m^2_{\tilde{\tau}_1}$ and seems to be favorable to the constraint of $m^2_{\tilde{\tau}_1}$. However, as $m_{D1}^2$ decreases, $m^2_{\tilde{\tau}_L}$ is decreased and below a critical value of $m_{D1}^2$ $m^2_{\tilde{\tau}_L}$ becomes dominant to $m^2_{\tilde{\tau}_1}$. A similar effect of the $D$-term contribution has been discussed within the framework of $SO(10)$ GUTs \cite{kks}. On the other hand, a positive value of $m_{D2}^2$ increases both $m^2_{\tilde{\tau}_R}$ and $m^2_{\tilde{\tau}_L}$. Thus, a positive value of $m_{D2}^2$ is favorable to increase $m^2_{\tilde{\tau}_1}$. The effects of $m_{D2}^2$ to $m_{H1}^2$ and $m_{H2}^2$ are same. Hence, $m_{D2}^2$ has little effect on electroweak symmetry breaking. These results are shown in Figs.3 and Figs. 4 for $\theta = \pi/2$. Figs. 3 show the $m_{D2}^2$ effects. A large value of $m_{D2}^2$ reduces the region excluded by the constraint $m^2_{\tilde{\tau}_1} >0$, while it has little effect on electroweak symmetry breaking. The region with the neutralino LSP appears and it becomes wider as $m_{D2}^2$ increases. For example, in the case of $m_{D2}^2=1.0$ (TeV)$^2$, the region with the neutralino LSP has the largest value $m_{3/2}=3.1$ TeV when $\alpha(T+\bar{T})=0.8$. In this case the lightest neutral Higgs boson mass $m_{h^0}$ is 151 GeV. Here we use the two loop Higgs mass $m_{h^0}$\cite{MSUSY,HIGS}. Similarly, we obtain $m_{h^0}=129$ GeV when $m_{3/2}=1.0$ TeV and $\alpha(T+\bar{T})=0.8$. This value is stable between $m_{3/2}=0.4$ TeV and 1.0 TeV. Figs. 4 show the $m_{D1}^2$ effects. They show that the region excluded by the electroweak symmetry breaking conditions becomes more narrow and the region with $m^2_{\tilde{\tau}_1} <0$ becomes wider as $m_{D1}^2$ decreases. In suitable values of $m_{D1}^2$, we have the region with the neutralino LSP. Figs. 4 show the critical value is around $m_{D1}^2=-2.0$ (TeV)$^2$, where the neutralino LSP region is widest. \subsubsection{Small $\tan \beta$ scenario} It is also possible that we obtain a smaller gauge group, e.g. $G_{SM}$ or $SU(5)$ at the GUT scale by non-standard embedding including Wilson line breaking. In this case, the assumption of the top-bottom-tau Yukawa coupling unification can be relaxed and a smaller value of $\tan \beta$ can be realized. We assume that below the GUT scale the matter content is exactly same as the MSSM. Here, we vary tan$\beta$ holding eqs.(\ref{Mg}),(\ref{ms}),(\ref{A}) as the initial conditions of RGEs at the GUT scale and fixing $M_t=175$ GeV. Figs. 5 show that the allowed parameter region for tan$\beta$. Fig. 5a shows that for a small value of $|\sin \theta|$ we have no excluded region at $\tan \beta < 50$. However, for large values of $|\sin \theta|$ and $2 \alpha Re(T)$, both the regions with $m^2_{\tilde{\tau}_1} <0$ and the stau LSP remain even when $\tan \beta $ is small as shown in Fig. 5b. \section{M-theory with 5-brane} \subsection{soft parameters} In general, non-standard embedding includes 5-branes, that is, 4-dimensional effective theory includes 5-brane moduli fields $Z^n$. It is possible that $Z^n$ also contribute to SUSY breaking. Thus, we consider such effects for generic case. If the 5-branes exist between the boundaries in M-theory, the K$\ddot{\mbox{a}}$hler potential and the gauge kinetic functions are modified. The relevant part of the moduli K$\ddot{\mbox{a}}$hler potential $K_{mod}$ and the gauge kinetic functions of the observable and hidden sectors $f^{(1)}$ and $f^{(2)}$ with $N$ 5-branes take the following form \cite{5brane,5b-soft}, \begin{eqnarray} K_{mod}&=&-ln(S+\bar{S})-3ln(T+\bar{T})+K_5 \label{Kmod}, \\ f^{(1)}&=&S+\epsilon T\left(\beta^{(0)} +\sum_{n=1}^{N}(1-Z^n)^2\beta^{(n)}\right), \nonumber \\ f^{(2)}&=&S+\epsilon T\left(\beta^{(N+1)} +\sum_{n=1}^{N}(Z^n)^2\beta^{(n)}\right) \label{5gfunc}, \end{eqnarray} where $K_5$ is the K$\ddot{\mbox{a}}$hler potential for the 5-brane position moduli $Z^n$ and the expansion parameter $\epsilon=\left({\kappa}/{4\pi}\right)^{2/3}{2\pi^2\rho}/{V^{2/3}}$. In addition, $\beta^{(0)}$, $\beta^{(N+1)}$ are the instanton numbers on the observable sector boundary and the hidden sector boundary respectively, and $\beta^{(n)}$ $(n=1,\cdots,N)$ is a magnetic charge on the each 5-brane. They should satisfy the cohomology constraint, \begin{eqnarray} \sum_{n=0}^{N+1}\beta_i^{(n)}=0 \label{coho}, \end{eqnarray} which means physically that the net charges must be zero because the ``flux'' cannot get away anywhere in compact space. The part of the matter fields is obtained \begin{eqnarray} K_{mat}=\left(\frac{3}{(T+\bar{T})}+\frac{\epsilon\zeta} {(S+\bar{S})}\right)|\Phi|^2, \end{eqnarray} where \begin{eqnarray} \zeta=\beta^{(0)}+\sum_{n=1}^{N}(1-z^n)^2\beta^{(n)}\label{zeta}. \end{eqnarray} Following the generic formulae of refs.\cite{SW,ST-soft} again, we can derive the soft parameters. We assume $V_0=0$ again. The soft parameters are obtained, \begin{eqnarray} M_{1/2}&=&\frac{1} {(S+\bar{S})+\epsilon\zeta(T+\bar{T})} \Bigl[F^S+\epsilon\zeta F^T+\epsilon T F^n \zeta_n\Bigr] \\ m^2&=&m_{3/2}^2 -|F^S|^2\Bigl[\frac{1}{(S+\bar{S})^2} -\frac{9}{\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2}\Bigr] \nonumber \\ &-&|F^T|^2\Bigl[\frac{1}{(T+\bar{T})^2} -\frac{(\epsilon\zeta)^2} {\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2}\Bigr] \nonumber \\ &-&F^n\bar{F}^{\bar{m}}\Bigl[\frac{\epsilon\zeta_{n\bar{m}}(T+\bar{T})} {3(S+\bar{S})+\epsilon\zeta(T+\bar{T})} -\frac{\epsilon^2\zeta_n \zeta_{\bar{m}}(T+\bar{T})^2} {\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2}\Bigr] \nonumber \\ &+&3F^S\bar{F}^{\bar{T}}\frac{\epsilon\zeta} {\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2} +3F^S\bar{F}^{\bar{m}}\frac{\epsilon\zeta_{\bar{m}}(T+\bar{T})} {\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2} \nonumber \\ &-&F^T\bar{F}^{\bar{m}}\Bigl[\frac{\epsilon\zeta_{\bar{m}}} {3(S+\bar{S})+\epsilon\zeta(T+\bar{T})} -\frac{\epsilon^2\zeta\zeta_{\bar{m}}(T+\bar{T})} {\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2}\Bigr]+(c.c) \nonumber \\ &{}& \\ A&=&F^S\Bigl[\frac{2}{S+\bar{S}} -\frac{9}{3(S+\bar{S})+\epsilon\zeta(T+\bar{T})}\Bigr] \nonumber \\ &+&F^T\frac{\epsilon\zeta}{3(S+\bar{S})+\epsilon\zeta(T+\bar{T})} \nonumber \\ &+&F^n\Bigl[K_{5,n}-\frac{3\epsilon\zeta_n(T+\bar{T})} {3(S+\bar{S})+\epsilon\zeta(T+\bar{T})}\Bigr], \end{eqnarray} where $K_{5,n}=\partial K_5/\partial Z^n$, $\zeta_{n}=\partial \zeta/\partial z^n$ and $\zeta_{nm}=\partial^2 \zeta/(\partial z^n \partial z^m)$. We have normalized the soft scalar mass as, \begin{eqnarray} m_{IJ}^2=m^2{\partial^2 K_{mat} \over |\partial \Phi |^2}, \end{eqnarray} using the total K\"ahler metric. That is different from ref.\cite{5b-soft}, where the soft scalar mass is normalized as, \begin{eqnarray} m_{IJ}^2=m^2{3 \over (T +\bar T)^2}. \end{eqnarray} In addition, here we have kept all terms, although in ref.\cite{5b-soft} only the linear terms of $\varepsilon$ have been kept. Here we ignore CP phases, i.e. $F=\bar{F}$. For simplicity, we assume that there is only one relevant 5-brane moduli $Z$ and its K\"ahler potential is a function of only $(Z+\bar Z)$, i.e., \begin{eqnarray} K_5=K_5(Z+\bar{Z}). \label{asmp1} \end{eqnarray} The gravitino mass is obtained, \begin{eqnarray} m_{3/2}^2=\frac{|F^S|^2}{3(S+\bar{S})^2}+\frac{|F^T|^2}{(T+\bar{T})^2} +\frac{1}{3}K_{5,Z\bar{Z}}|F^Z|^2\label{5mgra}. \end{eqnarray} Now we can parameterize each $F$-term as follows \begin{eqnarray} F^S&=&\sqrt{3}m_{3/2}(S+\bar{S})\sin\theta \sin\phi, \nonumber \\ F^T&=&m_{3/2}(T+\bar{T})\cos\theta \sin\phi, \nonumber \\ F^Z&=&\sqrt{\frac{3}{K_{5,Z\bar Z}}}m_{3/2}\cos\phi. \end{eqnarray} Still now we have several unknown parameters and most of them are due to detailed information of the 5-brane. Thus, we fix $z$ and $\beta^{(1)}$, e.g. \begin{eqnarray} z=\frac{1}{2},\hspace{1cm} \beta^{(1)}=\frac{4}{3}\beta^{(0)} \label{asmp2}. \end{eqnarray} Note that both the instanton number $\beta^{(0)}$ and the intersection number $\beta^{(1)}$ are integers. Eq.(\ref{asmp2}) leads to the following simple relations with eq.(\ref{zeta}) as \begin{eqnarray} \zeta|_{z={1}/{2}}=-\zeta_{,z}|_{z={1}/{2}} =\frac{1}{2}\zeta_{,zz}|_{z={1}/{2}} =\beta^{(1)} \label{relazeta}. \end{eqnarray} In this case we can rewrite the soft parameters, \begin{eqnarray} M_{1/2}&=&\frac{m_{{3}/{2}}}{(S+\bar{S})+\epsilon\zeta(T+\bar{T})} \left\{ \sqrt{3}(S+\bar{S}) \sin\theta \sin\phi \right. \nonumber \\ &+&\epsilon\zeta(T+\bar{T})\cos\theta \sin\phi -\left.\frac{\epsilon\zeta(T+\bar{T})}{2}\sqrt{\frac{3} {K_{5,Z\bar Z}}}\cos\phi \right\}, \nonumber\\ m^2&=&m_{3/2}^2 \left\{ 1-3\Bigl[1-\frac{9(S+\bar{S})^2}{\Bigl(3(S+\bar{S}) +\epsilon\zeta(T+\bar{T})\Bigr)^2}\Bigr]\sin^2\theta \sin^2\phi \right. \nonumber \\ &-&\Bigl[1-\Bigl(\frac{\epsilon\zeta(T+\bar{T})} {(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})}\Bigr)^2\Bigr] \cos^2\theta \sin^2\phi \nonumber \\ &-&\frac{3\epsilon\zeta(T+\bar{T})}{K_{5,Z \bar Z} \Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)} \Bigl[2-\frac{\epsilon\zeta(T+\bar{T})} {(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})}\Bigr]\cos^2\phi \nonumber \\ &+&\frac{6\sqrt{3}(S+\bar{S})\epsilon\zeta(T+\bar{T})} {\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2} \sin\theta \cos\theta \sin\phi \nonumber \\ &-&\frac{6\sqrt{3}(S+\bar{S})\epsilon\zeta(T+\bar{T})} {\Bigl(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})\Bigr)^2} \sqrt{\frac{3}{K_{5,Z \bar Z}}}\sin\theta \sin\phi \cos\phi \nonumber \\ &+&\frac{2\epsilon\zeta(T+\bar{T})}{3(S+\bar{S})+\epsilon\zeta(T+\bar{T})} \Bigl[1-\frac{\epsilon\zeta(T+\bar{T})} {(3(S+\bar{S})+\epsilon\zeta(T+\bar{T})}\Bigr]\nonumber \\ &\times& \left. \sqrt{\frac{3}{K_{5,Z \bar Z}}}\cos\theta \sin\phi \cos\phi \right\}, \nonumber \\ &{}& \nonumber\\ A&=&m_{3/2} \left\{ \sqrt{3}\Bigl[2 -\frac{9(S+\bar{S})}{3(S+\bar{S}) +\epsilon\zeta(T+\bar{T})}\Bigr]\sin\theta \sin\phi \right. \nonumber \\ &-&\frac{3\epsilon\zeta(T+\bar{T})} {3(S+\bar{S})+\epsilon\zeta(T+\bar{T})}\cos\theta \sin\phi \nonumber \\ &+& \left. \Bigl[K_{5,Z}+\frac{3\epsilon\zeta(T+\bar{T})} {3(S+\bar{S})+\epsilon\zeta(T+\bar{T})}\Bigr] \sqrt{\frac{3}{K_{5,Z \bar Z}}}\cos\phi \right\}, \label{5soft} \end{eqnarray} where we have assumed $T=(T+\bar{T})/2$. In addition, from the gauge kinetic functions (\ref{5gfunc}) we obtain constraints for values of moduli $S$ and $T$ as, \begin{eqnarray} 2Re(f^{(1)})=(S+\bar{S})+\epsilon\zeta(T+\bar{T})=2g_{GUT}^{-2}\simeq 4, \label{5const} \end{eqnarray} and \begin{eqnarray} 2Re(f^{(2)})=(S+\bar{S})-\frac{3}{2}\epsilon\zeta(T+\bar{T})>0.\label{5const2} \end{eqnarray} Here we use eqs.(\ref{asmp2}), (\ref{relazeta}), (\ref{coho}), (\ref{5const}) and (\ref{5const2}), so that we find the parameter region as $0< \epsilon\zeta(T+\bar{T}) < 8/5$. That is almost comparable to the region $0< \alpha(T+\bar{T}) < 2$ in the case without 5-branes. \subsection{Electroweak symmetry breaking and mass spectrum} Using eq.(\ref{5soft}), we analyze the 5-brane effects on electroweak symmetry breaking and the mass spectrum. Eq.(\ref{5soft}) has total 6 free parameters, that is, $m_{3/2}$, $\epsilon\zeta(T+\bar{T})$, $\theta$, $\phi$, and $K_{5,Z \bar Z}$, $K_{5,Z}$. Here we use the same notation for $\epsilon\zeta(T+\bar{T})$ as the case without 5-brane, i.e., $\alpha (T+\bar{T}) \equiv \epsilon\zeta(T+\bar{T})$. Note that $K_{5,Z \bar Z}$ is positive definite value due to eq.(\ref{5mgra}), (\ref{asmp1}), but $K_{5,Z}$ is any real number. The parameter $K_{5,Z}$ appears only in the $A$-parameter and $K_{5,Z}$-dependence of $A$ is easy to understand, that is, $A/\cos \phi$ increases linearly as $K_{5,Z}$ increases. Now, let us concentrate ourselves to the region with $|F^S|>>|F^T|$, that is, $\sin \theta \approx 1$. In this region, eq.(\ref{5soft}) is reduced to \begin{eqnarray} M_{1/2}&=&\frac{m_{\frac{3}{2}}}{4} \left\{ \sqrt{3}(4-\alpha(T+\bar{T}))\sin\phi -\frac{\alpha(T+\bar{T})}{2}\sqrt{\frac{3}{K_{5,Z \bar Z}}} \cos\phi \right\}, \nonumber \\ m^2&=&m_{3/2}^2 \left\{ 1-3\Bigl[1-\frac{9(4-\alpha(T+\bar{T}))^2}{\Bigl(3(S+\bar{S}) +\alpha(T+\bar{T})\Bigr)^2}\Bigr]\sin^2\phi \right. \nonumber \\ &-&\frac{3\alpha(T+\bar{T})}{K_{5,Z \bar Z} \Bigl(3(S+\bar{S})+\alpha(T+\bar{T})\Bigr)} \Bigl[2-\frac{\alpha(T+\bar{T})} {(3(S+\bar{S})+\alpha(T+\bar{T}))}\Bigr]\cos^2\phi \nonumber \\ &-&\left. \frac{6\sqrt{3}(4-\alpha(T+\bar{T}))\alpha(T+\bar{T})} {\Bigl(3(S+\bar{S})+\alpha(T+\bar{T})\Bigr)^2} \sqrt{\frac{3}{K_{5, Z \bar Z}}}\sin\phi \cos\phi \right\}, \nonumber \\ &{}& \nonumber \\ A&=&m_{3/2} \left\{ \sqrt{3}\Bigl[2 -\frac{9(4-\alpha(T+\bar{T}))}{3(S+\bar{S}) +\alpha(T+\bar{T})}\Bigr]\sin\phi \right. \nonumber \\ &+& \left. \Bigl[K_{5,Z}+\frac{3\alpha(T+\bar{T})} {3(S+\bar{S})+\alpha(T+\bar{T})}\Bigr] \sqrt{\frac{3}{K_{5,Z \bar Z}}}\cos\phi \right\}. \label{dilatonsoft} \end{eqnarray} To compare with the results in section 2 without 5-brane, here we take $\tan \beta =53$. Actually, in the case with such large $\tan \beta$ we have stronger constraints. At first, we consider the case with $K_{5,Z \bar Z}=K_{5,Z}=1$. Fig. 6 shows $M_{1/2}$, $m$ and $A$ for $\phi =\pi/4$ as well as $m_{3/2}=2.0$ TeV, $\theta =\pi/2$, and $K_{5,Z\bar Z}=K_{5,Z}=1$. Obviously, the ratios, $A/M_{1/2}$ and $m/M_{1/2}$, are different from eq.(\ref{sumrule}). The $A$-parameter is always positive when $\phi =\pi/4$. For small $\alpha(T+\bar{T})$, the dominant terms are the first term of $M_{1/2}$, and the second term of $m^2$. The 5-brane SUSY breaking effect $F^Z$ within $0<\sin\phi<1$ reduces $M_{1/2}$, but increases $m^2$. That is obvious from the comparison between Fig. 6 and eq.(\ref{sumrule}). Therefore we obtain $m_{\tilde{\tau}_1}^2>m_{\tilde{\chi}_1^0}^2 \sim M_1^2$ at small $\alpha(T+\bar{T})$ and the lightest neutralino is almost Bino-like. When $\alpha(T+\bar{T})$ become larger, the third and forth term of $m^2$ become more sizable and $m^2$ itself is decreased. Thus the excluded region by constraint $m_{\tilde{\tau}_1}^2>0$ is larger than the case without the 5-brane effect. Such behavior on electroweak symmetry breaking and the constraint $m_{\tilde{\tau}_1}^2>0$ is shown in Fig. 7a for $\theta =\pi/2$, $\phi =\pi/4$ and $K_{5,Z \bar Z}=K_{5,Z}=1$. In the allowed region, the lightest neutralino is almost Bino-like. Fig. 7b shows the case with $\phi =\pi/5$. In the 5-brane dominant case, $\cos \phi \to 1$, most of the region is excluded. Figs. 8a and 8b show $\alpha(T+\bar T)$-dependence and $\phi$-dependence of representative superparticle masses, e.g. the lightest neutralino, chargino, stop and stau. The $\phi$-dependence is rather larger than $\alpha(T+\bar T)$-dependence. When we vary $\phi$, the lightest Higgs mass change $132$ GeV $< m_{h^0} < 138$ GeV in the allowed region for $m_{3/2}= 1$ TeV and $\theta =\pi/2$. The lightest Higgs mass is rather stable for change of $\alpha(T+\bar T)$. Next, we calculate $K_{5,Z \bar Z}$ contribution fixing $K_{5,Z} =1$. Fig. 9 shows $M_{1/2}$, $m$ and $|A|$ for $m_{3/2}=2.0$ TeV, $\theta =\pi/2$, $\phi =\pi/4$, $2\alpha Re(T)=1.0$ and $K_{5,Z}=1$. Obviously, the $A/M_{1/2}$ and $m/M_{1/2}$, are different from eq.(\ref{sumrule}), again. The $A$-parameter changes from positive to negative around $K_{5,Z \bar Z} =2.5$ for $\phi =\pi/4$ and $2\alpha Re(T)=1.0$. The terms of the 5-brane contribution of $M_{1/2}$ and $m^2$ are proportional to $\sqrt{3/K_{5,Z \bar Z}}$ or $3/K_{5,Z \bar Z}$. When $K_{5,Z \bar Z}$ increases, then these terms are suppressed and these soft parameters become close to those without 5-brane except for the existence of the ``reducing'' factor $\sin\phi$. Therefore the effects of $\sin\phi$ is favorable for the $m_{\tilde{\tau}_1}^2>0$ condition and it leads to $m_{\tilde{\tau}_1}^2>m_{\tilde{\chi}_1^0}^2$. Fig. 10 shows such $K_{5,Z \bar Z}$-dependence of the regions excluded by the electroweak symmetry breaking and the constraint $m_{\tilde{\tau}_1}^2>0$ as well as the stau LSP region. Actually, large $K_{5,Z \bar Z}$ is useful to increase $m_{\tilde{\tau}_1}^2$, but it is not favorable for successful electroweak symmetry breaking. Finally, we show $K_{5,Z}$-dependence, which has the linear effect only on $A$. Figs. 11 show the $K_{5,Z}$ effect on the electroweak symmetry breaking, the constraint $m_{\tilde{\tau}_1}^2>0$ and the LSP for $m_{3/2}=2.0$ TeV, $\phi = \pi/4$ and $K_{5,Z \bar Z}=1$. As $K_{5,Z}$ increases, the electroweak symmetry breaking condition is relaxed, but the constraint $m_{\tilde{\tau}_1}^2>0$ becomes severe. Actually, the region with unsuccessful electroweak symmetry breaking disappears at $K_{5,Z}>1.4$, but at $K_{5,Z}>3.2$ all the region is excluded because of $m_{\tilde{\tau}_1}^2 <0$. Fig. 12 shows $K_{5,Z}$-dependence of representative superparticle masses. The lightest Higgs mass is rather stable against change of $K_{5,Z}$. \section{Conclusions} We have studied soft SUSY breaking parameters in M-theory with and without 5-brane moduli effects. We investigate successful electroweak symmetry breaking and the constraint $m_{\tilde{\tau}_1}^2>0$. In the allowed regions, we have shown mass spectra. We have considered the large $\tan \beta$ scenario, because these constraints are severe for large $\tan \beta$. In the case without 5-brane effects, the electroweak symmetry breaking conditions are relaxed in the strong coupling region $\alpha(T +\bar T) > 1.5$. In such region, however, the constraint $m_{\tilde{\tau}_1}^2>0$ becomes strong in particular for large $\theta$. In any allowed region, the LSP is the stau. The positive $D$-term contribution $m_{D2}^2$ can relax the constraint $m_{\tilde{\tau}_1}^2>0$ and lead to the region with the neutralino LSP. The negative $D$-term contribution $m_{D1}^2$ can relax the electroweak symmetry breaking conditions. In the suitable value of $m_{D1}^2$ we have the widest region with the neutralino LSP. For small $\tan \beta$, in particular $\tan \beta < 50$, the electroweak symmetry breaking is realized easily. However, for large $\alpha(T +\bar T)$ and large $\theta$, the region with $m_{\tilde{\tau}_1}^2 <0$ and the region with the stau LSP remain. The 5-brane effect can relax the electroweak symmetry breaking condition, but it makes the $m_{\tilde{\tau}_1}^2>0$ constraint severe. Soft terms in M-theory with 5-branes include more free parameters. Thus, we will leave to future work systematic study of the whole parameter space on other phenomenological aspects, e.g. cosmological implications and the $b \to s \gamma$ decay \cite{bs}, which leads to anther constraint for large $\tan \beta$ and lighter superparticles in the universal case. \section*{Acknowledgement} The authors would like to thank K.~Enqvist and S.-J. Rey for useful discussions. This work was partially supported by the Academy of Finland (no. 44129).
\section{Introduction} \label{sec:1} Analytical work in recent years has produced a number of systems of evolution equations which are equivalent to the Einstein equations at the constraint manifold, and which have a well posed initial value formulation~\cite{bona-masso92,frittelli-reula94,massoletter,york,frittelliletter,helmut96,frittelli-reula99}. What motivates interest in this type of result is a general understanding (see for instance~\cite{kreissbook}) that explicit well-posedness would be relevant in implementing consistent and stable numerical algorithms to integrate blackhole spacetimes. The well-posed schemes for which a numerical code has been implemented appear not to exibit significant improvements over other methods, there being several factors relevant to numerical implementation which play a significant role. What is puzzling, however, is that, on the other hand, there have been numerical simulations with apparently better behavior, but which are based on systems which do not seem to have the well-posed character. One preponderant feature of these numerically more robust schemes is that they are built on a decomposition of the intrinsic metric into a metric of unit determinant and the determinant itself, and of the extrinsic curvature into trace and trace-free part. With slight variations, this way of evolving the 3+1 Einstein equations has been considered by~\cite{kurki-laguna-matzner93,shibata}. Quite recently, this form has been shown to possess striking computational advantages over the standard form~\cite{BS99}. We refer to this general scheme as a conformally-decomposed formulation of the 3+1 Einstein equations. It is difficult to explain the success of these systems as opposed to the well-posed evolution schemes, or even to the standard (ADM) evolution schemes. The relative sizes of the fields of well-posed systems are roughly the same for different spectral frequencies in a Fourier representation, which helps explain the stability of the system via numerical analysis. However, in the conformally-decomposed systems this does not happen in general (for standard norms), as it does not happen for the standard ADM system, thus making it more difficult to justify their relative better behavior. We can speculate on two features that can possibly bear relevance to well behaved numerical evolution. One feature is that good behavior in evolution is related to constraint violations. If the system preserves more accurately the constraints, then the evolution remains closer to the constraint submanifold, which contains the physical solutions. Outside this submanifold the solutions are unphysical; thus, there is no compelling reason to rule out fast growths for seemingly tame initial data for unphysical solutions. Thus, we suggest that controlling the constraint violations may lead to well-behaved numerical evolution. In this respect, it has been shown~\cite{lambdavar}, that (at least in the linearized case) there are well-posed modifications of the Einstein equations outside the constraint submanifold which make that submanifold an attractor, thus improving the chances of building numerical codes with better behaved constraint propagation. Another possible cause of concern for generating numerical instabilities is the nature of the boundary conditions which are ussually imposed. The initial-boundary value problem for the Einstein equations has not been studied for the systems used in numerical simulations and where instabilities have been found (see nevertheless~\cite{helmutnagy} for a complete theory of boundary values for Einstein's equations in conformal frame variables, and~\cite{stewart98} for a linearized study), thus the set of boundary conditions for which the constraint equations are satisfied is in general not known. In dealing with this problem, establishing well-posedness for the Cauchy problem is a necessary first step. The other feature which could give rise to numerical instabilities is the relative sizes of the ``longitudinal'' and the ``radiative'' modes in general relativity. In all non-trivial asymptotically flat solutions (either vacuum or with matter satisfying the appropiate energy conditions) the positivity of the mass implies the existence of longitudinal modes, and there are many astrophysically relevant cases where there is an approximate local notion of longitudinal versus transverse modes, and where the former are several orders of magnitude bigger that the latter. If they are not properly separated in the numerical algorithms, the errors caused by finite differencing might be of the order of the ``radiative'' modes, and bad behavior can be expected. The separation of the conformal freedom in the more successful codes can perhaps be thought of as a way of dealing with this issue, or at least isolating it. In this work, we focus on this latter aspect. A technique for taking advantage of the conformal factor to partially decouple the ``longitudinal'' and ``transversal'' modes was used to obtain results on the Newtonian limit of general relativity~\cite{frittelli-reula94}. In that case the conformal field was fixed via an elliptic equation, decoupling in this way the more prominent Newtonian potential to first order from the radiative degrees of freedom. Further studies on this problem would be critical to obtain realistic simulations of most astrophysically relevant problems. Here we construct 3-parameter families of first-order well-posed systems wich share some of the properties of the more successful systems, such as the conformal decomposition of the fundamental fields, in the hope that their study would help understand what is causing them to behave better than others. In Section~\ref{sec:2} we apply techniques similar to those we used in~\cite{frittelliletter,frittelli-reula99} in order to obtain versions of the 3+1 equations that are conformally-decomposed but which are well posed. Additionally, we calculate the structure of characteristics and show that for a open region in parameter space the resulting equations are metric-causal, namely they have all propagation cones inside or coincident with the light cone. There is even a one parameter subfamily for which propagation is either along the light cone or normal to the slices. Furthermore we show that the constraints are propagated by these well-posed evolution equations. As oposed to~\cite{brugman}, where also analytical studies of systems with this decomposition have been done, in this work the trace of the extrisic curvature and the determinant of the intrinsic metric are considered dynamical variables and are evolved jointly with the rest of the system. \section{System II}\label{sec:2} The conformally-decomposed system that we take as starting point has appeared in~\cite{BS99}, and is a variation of the system used by Shibata and Nakamura~\cite{shibata}. It is a system of 15 equations for 15 variables $(\phi, K, \tilde{\gamma}_{ij}, \tilde{A}_{ij}, \tilde{\Gamma}^i)$, and is referred to as System II in~\cite{BS99}, to distinguish it from the standard 3+1 Einstein equations~\cite{york79}. These variables are related to the intrinsic metric $\gamma_{ij}$ and extrinsic curvature $K_{ij}$ as follows: \begin{mathletters} \begin{eqnarray} e^{4\phi} &=& \det(\gamma_{ij})^{1/3}\\ \tilde{\gamma}_{ij} & = & e^{-4\phi}\gamma_{ij}\\ K &=& \gamma^{ij}K_{ij}\\ \tilde{A}_{ij} &=& e^{-4\phi}\left( K_{ij} - \frac13 \gamma_{ij}K\right) \\ \tilde{\Gamma}^i &=& -\tilde{\gamma}^{ij},_j \end{eqnarray} \end{mathletters} where $\tilde{\gamma}^{ij}$ is the inverse of $\tilde{\gamma}_{ij}$. The Einstein equations in terms of these variables are equivalent to the following: \begin{mathletters}\label{bs} \begin{eqnarray} \frac{d}{dt}\phi &=& -\frac16 \alpha K\\ \frac{d}{dt}\tilde{\gamma}_{ij}& = & -2\alpha\tilde{A}_{ij}\\ \frac{d}{dt}K &=& -\gamma^{ij}D_iD_j\alpha +\alpha\left(\tilde{A}_{ij}\tilde{A}^{ij} +\frac13 K^2 \right) +\frac12\alpha(\rho+S)\\ \frac{d}{dt}\tilde{A}_{ij} &=& e^{-4\phi}\left( -(D_iD_j\alpha)^{TF} +\alpha(R_{ij}^{TF} - S_{ij}^{TF}) \right) +\alpha(K\tilde{A}_{ij}-2\tilde{A}_{il}\tilde{A}^l{}_j) \\ \frac{\partial}{\partial t}\tilde{\Gamma}^i &=& -2\tilde{A}^{ij}\alpha_j +2\alpha\left(\tilde{\Gamma}^i_{jk}\tilde{A}^{kj} -\frac23\tilde{\gamma}^{ij}K,_j -\tilde{\gamma}^{ij}S_j +6\tilde{A}^{ij}\phi,_j \right) \nonumber\\ && -\frac{\partial}{\partial x^j} \left( \beta^l\tilde{\gamma}^{ij},_l -2\tilde{\gamma}^{m(j}\beta^{i)},_m +\frac23\tilde{\gamma}^{ij}\beta^l,_l \right). \end{eqnarray} Here $\alpha$ is the lapse function, $\beta^i$ is the shift vector, and $\tilde{\Gamma}^i_{jk}$ are the connection coefficients of $\tilde{\gamma}_{ij}$. The superscript $^{TF}$ denotes trace-free part, e.g. $R_{ij}^{TF} = R_{ij} - \gamma_{ij}R/3$. Indices are raised and lowered with $\tilde{\gamma}^{ij}$ and its inverse. We use the shorthand notation \begin{equation} \frac{d}{dt} \equiv \frac{\partial}{\partial t} - \pounds_\beta \end{equation} where $\pounds_\beta$ is the Lie derivative along $\beta^i$. We have, as well, \begin{eqnarray} R^{TF}_{ij} &=& R_{ij}-\frac13 \gamma^{ij}\gamma^{kl}R_{kl}\\ R_{ij}&=&\tilde{R}_{ij}+R^{\phi}_{ij}\\ R^{\phi}_{ij}&=& -2\tilde{D}_i\tilde{D}_j\phi -2\tilde{\gamma}_{ij} \tilde{\gamma}^{kl}\tilde{D}_k\tilde{D}_l\phi +4 \tilde{D}_i\phi\tilde{D}_j\phi -4\tilde{\gamma}_{ij}\tilde{\gamma}^{kl} \tilde{D}_k\phi\tilde{D}_l\phi\\ \tilde{R}_{ij} &=& -\frac12\tilde{\gamma}^{kl}\tilde{\gamma}_{ij,kl} +\tilde{\gamma}_{k(i}\tilde{\Gamma}^k{},_{j)} \nonumber\\ && +\tilde{\Gamma}^k\tilde{\Gamma}_{(ij)k} +\tilde{\gamma}^{lm}\left( 2\tilde{\Gamma}^k{}_{l(i}\tilde{\Gamma}_{j)km} +\tilde{\Gamma}^k{}_{il}\tilde{\Gamma}_{kmj} \right) \\ \tilde{\Gamma}^k{}_{ij} &=& \frac12\tilde{\gamma}^{kl}\left( \tilde{\gamma}_{il,j} +\tilde{\gamma}_{jl,i} -\tilde{\gamma}_{ij,l} \right). \end{eqnarray} \end{mathletters} This system is first-order in time and second-order in space, thus it is of second order overall. We show how System II can be handled in order to be turned into a well-posed form. Firstly we reduce the system to a straightforward first order form, and subsequently we densitize the lapse and combine the constraints into the evolution equations. \subsection{System II reduced to first-order form} We define a set of 12 additional variables \begin{mathletters}\label{new} \begin{equation} V_{ijk} \equiv \tilde{\gamma}_{ij,k} -\frac35 \tilde{\gamma}_{k(i}\tilde{\gamma}_{j)n,s} \tilde{\gamma}^{ns} +\frac15 \tilde{\gamma}_{ij}\tilde{\gamma}_{kn,s} \tilde{\gamma}^{ns}, \end{equation} which is trace free in all its indices, namely: $V_{ijk}\tilde{\gamma}^{ij}=0$ and $V_{ijk}\tilde{\gamma}^{jk}= V_{ijk}\tilde{\gamma}^{ik}=0$, and another set of 3 additional variables \begin{equation} Q_i \equiv \phi,_i. \end{equation} \end{mathletters} Evolution equations for these new variables are obtained by taking a time derivative of (\ref{new}) and commuting time and spatial derivatives in the resulting right-hand sides. The complete system of equations is now \begin{mathletters}\label{firstbs} \begin{eqnarray} \frac{d}{dt}\phi &=& -\frac16 \alpha K\\ \frac{d}{dt}\tilde{\gamma}_{ij}& = & -2\alpha\tilde{A}_{ij}\\ \frac{d}{dt}K &=& -\gamma^{ij}D_iD_j\alpha +\alpha\left(\tilde{A}_{ij}\tilde{A}^{ij} +\frac13 K^2 \right) +\frac12\alpha(\rho+S)\\ \frac{d}{dt}\tilde{A}_{ij} &=& e^{-4\phi}\left( -(D_iD_j\alpha)^{TF} +\alpha(R_{ij}^{TF} - S_{ij}^{TF}) \right) +\alpha(K\tilde{A}_{ij} -2\tilde{A}_{il}\tilde{A}^l{}_j) \\ \dot{\tilde{\Gamma}^i} -\beta^l\tilde{\Gamma}^i,_l &=& -2\tilde{A}^{ij}\alpha,_j +2\alpha\left(\tilde{\Gamma}^i{}_{jk}\tilde{A}^{kj} -\frac23\tilde{\gamma}^{ij}K,_j -\tilde{\gamma}^{ij}S_j +6\tilde{A}^{ij}Q_j \right) -\beta^l,_j \tilde{\gamma}^{ij},_l \nonumber\\ && +\tilde{\Gamma}^m\beta^i,_m +\tilde{\gamma}^{mi},_j\beta^j,_m +2\tilde{\gamma}^{m(i}\beta^{j)},_{mj} +\frac23\tilde{\Gamma}^i\beta^l,_l +\frac23\tilde{\gamma}^{ij}\beta^l,_{lj} \\ \dot{V}_{ijk}-\beta^l V_{ijk,l} &=& -2\alpha \tilde{A}_{ij,k} +\frac65\alpha\tilde{\gamma}_{k(i} \tilde{A}_{j)m,n}\tilde{\gamma}^{mn} -\frac25\alpha\tilde{\gamma}_{ij} \tilde{A}_{km,n}\tilde{\gamma}^{mn} \nonumber\\ && +\beta^l,_kV_{ijl} +\beta^l,_i V_{ljk} +\beta^l,_j V_{ilk} -2\alpha,_k\tilde{A}_{ij} \\ && +\frac65\tilde{\gamma}_{k(i}\tilde{A}_{j)}^n \alpha,_n +\frac65\alpha\tilde{A}_{k(i}\tilde{\Gamma}_{j)} -\frac25\alpha\tilde{A}_{ij}\tilde{\Gamma}_k \\ && +\tilde{\gamma}_{l(i}\beta^l,_{j)k} -\frac35\beta^l,_{l(j}\tilde{\gamma}_{i)k} +\frac15\tilde{\gamma}_{ij}\beta^l,_{lk} \\ && -\beta^l,_{ns}\tilde{\gamma}^{ns} \left( \frac35\tilde{\gamma}_{k(i}\tilde{\gamma}_{j)l} -\frac15\tilde{\gamma}_{ij}\tilde{\gamma}_{kl} \right) \\ && +\left(2\alpha\tilde{A}^{ns} +\tilde{\gamma}^{ns}\beta^s,_m \right) \left(\frac35\tilde{\gamma}_{k(i} \tilde{\gamma}_{j)n,s} -\frac15\tilde{\gamma}_{ij} \tilde{\gamma}_{ks,n} \right) \\ \dot{Q}_i - \beta^l Q_{i,l} &=& -\frac16 \alpha K,_i +\beta^l,_i Q_l -\frac16 \alpha,_i K \end{eqnarray} \end{mathletters} where, as before, $R^{TF}_{ij} = R_{ij}-\frac13 \gamma^{ij} \gamma^{kl} R_{kl}$, with $R_{ij} = \tilde{R}_{ij} + R^{\phi}_{ij}$, and \begin{mathletters} \begin{eqnarray} R^{\phi}_{ij} &=& -2\tilde{D}_iQ_j -2\tilde{\gamma}_{ij} \tilde{\gamma}^{kl}\tilde{D}_kQ_l +4 Q_iQ_j -4\tilde{\gamma}_{ij}\tilde{\gamma}^{kl}Q_kQ_l,\\ \tilde{R}_{ij} &=& -\frac12 V_{ijk},^k +\frac{7}{10}\tilde{\gamma}_{k(i} \tilde{\Gamma}^k{},_{j)} +\frac{1}{10}\tilde{\gamma}_{ij} \tilde{\Gamma}^k,_k \nonumber\\ && -\frac{3}{10}\left( \tilde{\Gamma}^kV_{k(ij)} +\tilde{\Gamma}_{(i}V_{j)k}{}^k +\frac{9}{10} \tilde{\Gamma}_i\tilde{\Gamma}_j \right) \nonumber\\ && -\frac15\left( V_{ijk}\tilde{\Gamma}^k +\tilde{\gamma}_{ij} \tilde{\Gamma}^kV_{km}{}^m -\frac{1}{10} \tilde{\gamma}_{ij} \tilde{\Gamma}^k\tilde{\Gamma}_k \right) \nonumber\\ && +\tilde{\Gamma}^k\tilde{\Gamma}_{(ij)k} +2\tilde{\Gamma}^{kl}{}_{(i} \tilde{\Gamma}_{j)kl} +\tilde{\Gamma}^k{}_{il} \tilde{\Gamma}_{kj}{}^l, \\ \tilde{\Gamma}^k{}_{ij} &=& V^k{}_{(ij)} -\frac12 V_{ij}{}^k -\frac15\delta^k_{(i}\tilde{\Gamma}_{j)} +\frac25\tilde{\gamma}_{ij}\tilde{\Gamma}^k, \end{eqnarray} \end{mathletters} and indices are raised and lowered with $\tilde{\gamma}^{ij}$ and $\tilde{\gamma}_{ij}$ respectively. The derivatives of the form $\tilde{\gamma}_{ij,k}$ that appear in the right-hand sides of (\ref{firstbs}) must be interpreted simply as shorthands for combinations of the fields $\tilde{\Gamma}^i$ and $V_{ijk}$, via \begin{equation} \tilde{\gamma}_{ij,k} = V_{ijk} +\frac35\tilde{\gamma}_{k(i}\tilde{\Gamma}_{j)} -\frac15\tilde{\gamma}_{ij}\tilde{\Gamma}_k \end{equation} For this first-order system to be equivalent to the Einstein equations (in the sense that its set of solutions is the same as that of the Einstein equations), the following set of constraints must be imposed on the initial data (and are subsequently preserved by the evolution, as will be shown in the next section): \begin{mathletters}\label{const} \begin{eqnarray} {\cal H} &=& \gamma^{ij}R_{ij} - \tilde{A}_{ij} \tilde{A}^{ij} +\frac23 K^2 -2\rho \label{hamilt} \\ {\cal P}_i &=& \tilde{\gamma}^{jl}D_l\tilde{A}_{ij} -\frac23 D_i K +4Q_l\tilde{A}^l{}_i +\frac43 KQ_i -S_i \label{moment}\\ {\cal G}^i &=& \tilde{\Gamma}^i +\tilde{\gamma}^{ij},_j \label{gamma} \\ {\cal Q}_i &=& Q_i - \phi,_i \label{q}\\ {\cal V}_{ijk} &=& V_{ijk} - \tilde{\gamma}_{ij,k} +\frac35 \tilde{\gamma}_{k(i} \tilde{\gamma}_{j)n,s} \tilde{\gamma}^{ns} -\frac15 \tilde{\gamma}_{ij} \tilde{\gamma}_{kn,s} \tilde{\gamma}^{ns},\label{v} \end{eqnarray} where \begin{eqnarray} \gamma^{ij}R_{ij} &=& e^{-4\phi}\left( \tilde{\Gamma}^l,_l -8\tilde{D}^lQ_l -Q^lQ_l -\frac12 V_{ijl}V^{ijl} -\frac{15}{10}\tilde{\Gamma}^kV_{km}{}^m -\frac15 V^m{}_mk\tilde{\Gamma}^k \right. \\ && \left. -\frac{21}{100}\tilde{\Gamma}^k\tilde{\Gamma}_k +\tilde{\Gamma}^k\tilde{\Gamma}^m{}_{mk} +2\tilde{\Gamma}^{klm}\tilde{\Gamma}_{mkl} +\tilde{\Gamma}^{mkl}\tilde{\Gamma}_{mkl} \right). \end{eqnarray} \end{mathletters} Constraints (\ref{hamilt}) and (\ref{moment}) are the hamiltonian and momentum constraints of the 3+1 decomposition of the Einstein equations, written in our choice of variables. Constraints ({\ref{gamma}), (\ref{q}) and (\ref{v}) arise in turning the original second-order system into first order. In (\ref{firstbs}) and (\ref{const}), the derivative $D_l$ is the covariant derivative with respect to $\gamma_{ij}$, and is related to $\tilde D_l$ by undifferentiated terms: \begin{equation} \Gamma^k{}_{ij} = \tilde{\Gamma}^k{}_{ij} +2\left( Q_i\delta^k_j +Q_j\delta^k_i -Q_l\tilde{\gamma}^{kl}\tilde{\gamma}_{ij} \right) \end{equation} \subsection{Taking advantage of the available freedom} In this section we take advantage of two facts that have been used successfully in similar problems~\cite{choquet-ruggeri83,frittelliletter,frittelli-reula99}. Firstly, we densitize the lapse $\alpha$ (and in doing so we introduce a free function, referred as ``slicing density'' in~\cite{york98}): \begin{equation} \alpha=e^{4a\phi}\sigma \end{equation} Like the shift vector $\beta^i$, the lapse density $\sigma$ will be considered arbitrary but fixed, a source function independent of the dynamical fields. Secondly, the evolution equations can be combined with the constraints without altering the set of solutions. We add the scalar constraint with a factor $b$ to the evolution equation for $K$ and we add the vector constraint to the evolution equations for $\tilde{\Gamma}^i$ and $Q_i$, with factors of $c$ and $d$ respectively. In this manner we obtain a system of the form \begin{equation} \dot{u} = \mbox{\boldmath $A$}^i(u) \nabla_i u + B(u). \end{equation} A system of this form is known to be well posed if the matrix-valued vector $\mbox{\boldmath $A$}^i(u)$ admits a symmetrizer, namely, a positive definite, symmetric, bi-linear form \mbox{\boldmath $H$}, in the space of the fields $u$, whose product with $\mbox{\boldmath $A$}^i(u)$ yields a symmetric-bilinear-form-valued vector. Thus, in order to determine well-posedness, it suffices to consider the principal part of the system. In this case, the principal terms are \begin{mathletters}\label{combs} \begin{eqnarray} \dot{\phi} &=& \beta^l\phi,_l \\ \dot{\tilde{\gamma}}_{ij} &=& \beta^l\tilde{\gamma}_{ij,l} \\ \dot{K} &=& \beta^l K,_l -\alpha(4a+8b)e^{-4\phi}\tilde{\gamma}^{kl}Q_{k,l} +\alpha b e^{-4\phi}\tilde{\Gamma}^l{},_l \\ \dot{\tilde{A}}_{ij} &=& \beta^l\tilde{A}_{ij,l} +e^{-4\phi}\alpha\left( -\frac12 \tilde{\gamma}^{kl}V_{ijk,l} +\frac{7}{10}\left( \tilde{\gamma}_{k(i}\tilde{\Gamma}^k,_{j)} -\frac13\tilde{\gamma}_{ij}\tilde{\Gamma}^k,_k \right) \right) \nonumber\\ && -2(2a+1)e^{-4\phi}\alpha\left( Q_{(i,j)} -\frac13\tilde{\gamma}_{ij} \tilde{\gamma}^{kl}Q_{k,l} \right) \\ \dot{\tilde{\Gamma}}^i &=& \beta^l\tilde{\Gamma}^i,_l +\alpha c\tilde{A}^{il},_l -\frac23(c+2)\alpha \tilde{\gamma}^{il} K,_l \\ \dot{V}_{ijk} &=& \beta^l V_{ijk,l} -2\alpha \tilde{A}_{ij,k} +\frac65\alpha\tilde{\gamma}_{k(i} \tilde{A}_{j)m,n} \tilde{\gamma}^{mn} -\frac25\alpha\tilde{\gamma}_{ij} \tilde{A}_{km,n} \tilde{\gamma}^{mn} \\ \dot{Q}_i &=& \beta^l Q_{i,l} -\frac{\alpha}{6}(1+4d)K,_i +d \alpha\tilde{\gamma}^{jl}\tilde{A}_{ij,l}. \end{eqnarray} \end{mathletters} Our aim is to show that there exist choices of the numerical factors $a,b,c,d$ such that the system (\ref{combs}) is symmetrizable, therefore, well posed. We establish this result by defining a candidate symmetrizer \mbox{\boldmath$H$} given as \begin{eqnarray}\label{energy} \bar u \mbox{\boldmath$H$} u &=& \phi^2 + \delta^{ik}\delta^{jl}\tilde{\gamma}_{ij} \tilde{\gamma}_{kl} +n_1^2 e^{-4\phi} K^2 +\tilde{A}^{ij}\tilde{A}_{ij} +n_2^2 e^{-4\phi}\tilde{\Gamma}^i\tilde{\Gamma}_i +\frac{e^{-4\phi}}{4}V_{ijk}V^{ijk} \nonumber\\ && +n_3^2 e^{-4\phi} Q^iQ_i \end{eqnarray} where $n_1, n_2, n_3$ are any fixed real numbers different from zero and bounded, so that $C^{-1} \mbox{\boldmath$I$}\le \mbox{\boldmath$H$} \le C\mbox{\boldmath$I$}$ where $C$ is a positive constant and \mbox{\boldmath$I$} is the identity operator on the space of $u$ . We can easily arrange the values of $a,b,c,d$ so that symmetry of $\mbox{\boldmath$HA$}^i(u)$ is attained. To this effect, they must satisfy \begin{mathletters}\label{ens} \begin{eqnarray} \frac{7}{10} & = & n_2^2 c\\ -2(2a+1) & = & n_3^2 d\\ -(4a+8b) n_1^2 & = & -\frac16(1+4d)n_3^2\\ n_1^2 b & = & -\frac23 n_2^2 (c+2) \end{eqnarray} \end{mathletters} There is clearly plenty of freedom in the choice of $a,b,c,d$, since any choice that results in non-vanishing $n_1,n_2,n_3$ is allowed. The freedom is thus parametrized by the values of $n_1^2, n_2^2, n_3^2$, since these can take independent positive values. Thus our four parameters $a,b,c,d$ are not all independent, but there is a relationship between them that reduces the freedom to 3 independent parameters. We can solve (\ref{ens}) for $a,b,c,d$ in terms of $n_1, n_2, n_3$, which yields: \begin{mathletters}\label{parametric} \begin{eqnarray} a&=& \frac{9/5+8n_2^2+n_3^2/8}{(3n_1^2+2)}\\ b&=& -\frac{2}{3n_1^2}\left(\frac{7}{10}+2n_2^2\right)\\ c&=&\frac{7}{10n_2^2} \\ d&=&-\frac{2}{n_3^2} \left(2\frac{9/5+8n_2^2+n_3^2/8}{(3n_1^2+2)} +1\right) \end{eqnarray} \end{mathletters} It is clear from (\ref{parametric}) that $a$ and $c$ will take only strictly positive values, and $b$ and $d$ will take only strictly negative values, for all real values of $n_1, n_2, n_3$ different from zero. \subsection{Structure of characteristics} The system (\ref{combs}) is of the form \begin{equation} \mbox{\boldmath$A^a$}\frac{\partial u}{\partial x^a} = 0. \end{equation} The characteristic covectors are covectors $\xi_a=(\xi_i,-v)$ such that $\xi_i\xi_j\gamma^{ij}=1$ and such that \begin{equation}\label{det} \det(\mbox{\boldmath$A^a$}\xi_a)=0 \end{equation} The values of $v$ that satisfy (\ref{det}) for every direction $\xi_i$ are the characteristic speeds in that direction. In order to find these values we set up an eigenvalue problem for the principal symbol $\mbox{\boldmath$A^a$}\xi_a$ and find the null eigenvectors. The eigenvalue problem is \begin{mathletters}\label{eigencombs} \begin{eqnarray} n^a\xi_a\phi &=& 0 \label{eigencombs1}\\ n^a\xi_a\tilde{\gamma}_{ij}& = & 0 \label{eigencombs2}\\ n^a\xi_a K &=& -(4a+8b)e^{-4\phi}\xi^kQ_{k} + b e^{-4\phi}\xi_l\tilde{\Gamma}^l \label{eigencombs3}\\ n^a\xi_a\tilde{A}_{ij} &=& e^{-4\phi}\left( -\frac12 \xi^kV_{ijk} +\frac{7}{10}\left( \tilde{\gamma}_{k(i}\xi_{j)}\tilde{\Gamma}^k -\frac13\tilde{\gamma}_{ij}\xi_k\tilde{\Gamma}^k \right) \right) \nonumber\\ && -2(2a+1)e^{-4\phi}\left( \xi_{(j}Q_{i)} -\frac13\tilde{\gamma}_{ij}\xi^kQ_k \right) \label{eigencombs4}\\ n^a\xi_a\tilde{\Gamma}^i &=& c\xi_l\tilde{A}^{il} -\frac23(c +2) \xi^i K \label{eigencombs5}\\ n^a\xi_a V_{ijk} &=& -2\xi_k\tilde{A}_{ij} +\frac65\tilde{\gamma}_{k(i}\tilde{A}_{j)m}\xi^m -\frac25\tilde{\gamma}_{ij}\xi^m\tilde{A}_{km} \label{eigencombs6}\\ n^a\xi_a Q_i &=& -\frac16(1+4d)\xi_iK +d \xi^j\tilde{A}_{ij} \label{eigencombs7} \end{eqnarray} \end{mathletters} Clearly, $n^a\xi_a=0$ allows for 18 eigenvectors. This is because (\ref{eigencombs5}), (\ref{eigencombs6}) and (\ref{eigencombs7}) in this case constitute an overdetermined system of 18 homogeneous equations for 6 unknowns $(\tilde{A}_{ij}, K)$, with zero as the only solution, whereas (\ref{eigencombs3}) and (\ref{eigencombs4}) constitute a system of 6 equations for 18 variables, which leaves out 12 of the 18 fields $(V_{ijk}, \tilde{\Gamma}^i, Q)$ free. Lastly, (\ref{eigencombs1}) and (\ref{eigencombs2}) leave the 6 variables $(\tilde{\gamma}_{ij}, \phi)$ free. If we represent the eigenvectors in the form \begin{equation} (\tilde{\gamma}_{ij}, \phi, Q_i, \tilde{\Gamma}^{(L)}, \tilde{\Gamma}^{(T)}_i, V^{(L)}_{ij}, V^{(T)}_{ijk}, K, \tilde{A}^{(LL)}, \tilde{A}^{(LT)}_i, \tilde{A}^{(TT)}_{ij}) \end{equation} where $ \tilde{\Gamma}^{(L)}\!:= \tilde{\Gamma}^i\xi_i,\; \tilde{\Gamma}^{(T)}_i\!:=\tilde{\Gamma}_i- e^{-4\phi} \xi_i\tilde{\Gamma}^k\xi_k,\; V^{(L)}_{ij}\!:= V_{ijk}\xi^k, V^{(T)}_{ijk}\! := V_{ijk}- e^{-4\phi}\xi_k V_{ijl}\xi^l,\; \tilde{A}^{(LL)}\! := \tilde{A}^{ij}\xi_i\xi_j,\; \tilde{A}^{(LT)}_i\!:= \tilde{A}_{ij}\xi^j -e^{-4\phi}\xi_i \tilde{A}^{kl}\xi_k\xi_l ,\; \tilde{A}^{(TT)}_{ij}\! := \tilde{A}_{ij} -2e^{-4\phi}\xi_{(i}\tilde{A}_{j)l}\xi^l +e^{-8\phi}\xi_i\xi_j\tilde{A}^{kl}\xi_k\xi_l,\; $ then we have 5 eigenvectors corresponding to the five components of the conformal metric \begin{mathletters} \begin{equation} (\tilde{\gamma}_{ij}, 0,0,0,0,0,0,0,0,0); \end{equation} we have the determinant as an eigenfield \begin{equation} (0, \phi,0,0,0,0,0,0,0,0); \end{equation} we have 7 eigenvectors corresponding to the seven transverse components of $V_{ijk}$ \begin{equation} (0,0,0,0,0,0,V^{(T)}_{ijk},0,0,0); \end{equation} we have 3 eigenvectors corresponding essentially to the three components of $Q_i$ \begin{equation} (0,0,Q_i,\frac{4a+8b}{b}\xi^iQ_i,0, -2(2a+1)\xi_{(i}Q_{j)} -\frac13\tilde{\gamma}_{ij} \left(\frac{7(4a+8b)}{10b}-2(2a+1)\right)Q_l\xi^l,0,0,0,0); \end{equation} and we have 2 eigenvectors corresponding essentially to the two components of the transverse part of $\tilde{\Gamma}_i$ \begin{equation} (0,0,0,0,\tilde{\Gamma}^{(T)}_i, \frac{7}{10}\xi_{(i}\tilde{\Gamma}^{(T)}_{j)},0,0,0,0); \end{equation} \end{mathletters} If $n^a\xi_a\neq 0$, then $\tilde{\gamma}_{ij}=\phi=0$, and we can solve (\ref{eigencombs5}), (\ref{eigencombs6}) and (\ref{eigencombs7}) for $(V_{ijk}, \tilde{\Gamma}^i, Q)$ in terms of $\xi_a$ and $(\tilde{A}_{ij}, K)$. We can substitute $(V_{ijk}, \tilde{\Gamma}^i, Q)$ into (\ref{eigencombs3}) and (\ref{eigencombs4}), obtaining thus a system of 6 equations for the 6 variables $(\tilde{A}_{ij}, K)$ as follows \begin{mathletters}\label{red} \begin{eqnarray} 0 &=& K e^{4\phi}\left( (n^a\xi_a)^2 -\frac16(4a+8b)(1+4d) +\frac23 b(c+2) \right) +\xi\!\cdot\!\tilde{A}\!\cdot\!\xi \left( (4a+8b)d-bc \right) \label{red1}\\ 0 &=& \tilde{A}_{ij} e^{4\phi}\left(1-(n^a\xi_a)^2\right) -\frac{K}{3}\left( \frac{7}{5}(c+2) -(2a+1)(1+4d) \right) \left( \xi_i\xi_j -\frac13e^{4\phi}\tilde{\gamma}_{ij} \right) \nonumber\\ && +\left( \frac{7c}{10} -\frac35 -2d(2a+1) \right) \left( \xi^l\tilde{A}_{l(i}\xi_{j)} -\frac13\tilde{\gamma}_{ij} \xi\!\cdot\!\tilde{A}\!\cdot\!\xi \right) \label{red2} \end{eqnarray} \end{mathletters} where we have used the notation \begin{equation} \xi\!\cdot\!\tilde{A}\!\cdot\!\xi := \xi_i\tilde{A}^{ij}\xi_j. \end{equation} If $1-(n^a\xi_a)^2=0$, then $K=\xi\!\cdot\!\tilde{A}\!\cdot\!\xi=0$ by (\ref{red1}), which implies $\xi^l\tilde{A}_{li}=0$ by (\ref{red2}). However, two of the five components of $\tilde{A}_{ij}$ are thus free, which means that there are 4 eigenvectors, essentially labeled by the transverse components of $\tilde{A}_{ij}$. We have 2 eigenvectors for $n^a\xi_a=1$ \begin{mathletters} \begin{equation} (0,0,0,0,0, -2\tilde{A}^{(TT)}_{ij},0,0,0,0,\tilde{A}^{(TT)}_{ij}); \end{equation} and 2 eigenvectors for $n^a\xi_a=-1$ \begin{equation} (0,0,0,0,0, 2\tilde{A}^{(TT)}_{ij},0,0,0,0,\tilde{A}^{(TT)}_{ij}). \end{equation} \end{mathletters} If $1-(n^a\xi_a)^2\neq 0$, then contracting (\ref{red2}) with $\xi^i$ yields \begin{eqnarray} \label{rered} 0&=&e^{4\phi}\xi^l\tilde{A}_{lj} \left(1 +\frac12\left(\frac{7c}{10} -\frac35 -2d(2a+1) \right) -(n^a\xi_a)^2 \right) \nonumber\\ && -\frac29e^{4\phi} K\xi_j \left(\frac{7}{5}(c+2)- (2a+1)(1+4d) \right) \nonumber\\ && +\frac16\xi\!\cdot\!\tilde{A}\!\cdot\!\xi \;\xi_j \left(\frac{7}{10}c-\frac35 -2d(2a+1)\right) \end{eqnarray} Thus, if $1+\frac12(7c/10-3/5-2d(2a+1))-(n^a\xi_a)^2=0$, then $K=\xi\!\cdot\!\tilde{A}\!\cdot\!\xi=0$ by (\ref{red1}), which implies that (\ref{rered}) is identically satisfied, thus three out of the five equations (\ref{red2}) are identities, the remaining two determining two components of $\tilde{A}_{ij}$. Thus two of the five components of $\tilde{A}_{ij}$ are free, which means that there are 4 eigenvectors, essentially labeled by the two longitudinal-transverse components of $\tilde{A}_{ij}$. We have 2 eigenvectors for $n^a\xi_a = \sqrt{(3/5-7c/10-2d(2a+1))/2}$, namely \begin{mathletters} \begin{equation} (0,0,\frac{d}{C_1}\tilde{A}^{(LT)}_i,0, \frac{c}{C_1}\tilde{A}^{(LT)}_i, -\frac{4}{5 C_1}\xi_{(i}\tilde{A}^{(LT)}_{j)}, \frac{6}{5 C_1} \left(\tilde{\gamma}_{k(i}\tilde{A}^{(LT)}_{j)} -\xi_k\xi_{(i}\tilde{A}^{(LT)}_{j)} -\frac13\tilde{\gamma}_{ij}\tilde{A}^{(LT)}_k \right), 0,0,\tilde{A}^{(LT)}_i, 0), \end{equation} and 2 eigenvectors for $n^a\xi_a = -\sqrt{(3/5-7c/10-2d(2a+1))/2}$, namely \begin{equation} (0,0,-\frac{d}{C_1}\tilde{A}^{(LT)}_i,0, -\frac{c}{C_1}\tilde{A}^{(LT)}_i, \frac{4}{5 C_1}\xi_{(i}\tilde{A}^{(LT)}_{j)}, -\frac{6}{5 C_1} \left(\tilde{\gamma}_{k(i}\tilde{A}^{(LT)}_{j)} -\xi_k\xi_{(i}\tilde{A}^{(LT)}_{j)} -\frac13\tilde{\gamma}_{ij}\tilde{A}^{(LT)}_k \right), 0,0,\tilde{A}^{(LT)}_i, 0), \end{equation} \end{mathletters} where we have used the shorthand notation \begin{equation} C_1 :=\sqrt{(3/5-7c/10-2d(2a+1))/2} \end{equation} But if $1+\frac12(7c/10-3/5-2d(2a+1))-(n^a\xi_a)^2\neq 0$, then $\xi^l\tilde{A}_{lj}$ is determined by the values of $K$ and $\xi\!\cdot\!\tilde{A}\!\cdot\!\xi$ by (\ref{rered}), and if plugged back into (\ref{red2}) it follows that all the components of $\tilde{A}_{ij}$ are determined by $K$ and $\xi\!\cdot\!\tilde{A}\!\cdot\!\xi$. Therefore it is necessary that $K$ and $\xi\!\cdot\!\tilde{A}\!\cdot\!\xi$ be nonvanishing. Contracting (\ref{rered}) with $\xi^j$ we obtain \begin{eqnarray}\label{rerered} 0 &=& -\frac29 e^{4\phi}K\left(\frac{7}{5}(c+2) -(2a+1)(1+4d) \right) \nonumber\\ && +\xi\!\cdot\!\tilde{A}\!\cdot\!\xi \left(1 +\frac23\left(\frac{7}{10}c-\frac35-2(2a+1)d \right) -(n^a\xi_a)^2 \right) \end{eqnarray} Equations (\ref{red1}) and (\ref{rerered}) form a system of two homogeneous equations for $K$ and $\xi\!\cdot\!\tilde{A}\!\cdot\!\xi$. Thus, for $K$ and $\xi\!\cdot\!\tilde{A}\!\cdot\!\xi$ to be nonvanishing, it is necessary that the determinant of the system be zero. The determinant is \begin{equation}\label{chardet} \frac{1}{45} (-3(n^a\xi_a)^2+2a)(15(n^a\xi_a)^2-9+10bc-80bd+20d-7c) \end{equation} It can be seen that, because $a$ and $c$ are strictly positive and $b$ and $d$ are strictly negative, the four roots of the determinant are real. For the roots $n^a\xi_a$ of the determinant, we have \begin{mathletters} \begin{eqnarray} K &=& \frac92 \frac{1 -(n^a\xi_a)^2 +\frac23 \left(\frac{7c}{10}-\frac35-2d(2a+1) \right) } {\frac75(c+2)-(2a+1)(1+4d))} e^{-4\phi}\xi\!\cdot\!\tilde{A}\!\cdot\!\xi \\ Q_i &=& \left( d -\frac{3(1+4d)}{4} \frac{1 -(n^a\xi_a)^2 +\frac23 \left(\frac{7c}{10}-\frac35-2d(2a+1) \right) } {\frac75(c+2)-(2a+1)(1+4d))} \right) \frac{e^{-4\phi}}{n^a\xi_a} \xi_i\xi\!\cdot\!\tilde{A}\!\cdot\!\xi\\ \tilde{\Gamma}^{(L)} &=& \left( c -3(c+2)\frac{1 -(n^a\xi_a)^2 +\frac23 \left(\frac{7c}{10}-\frac35-2d(2a+1) \right) } {\frac75(c+2)-(2a+1)(1+4d))} \right)\frac{\xi\!\cdot\!\tilde{A}\!\cdot\!\xi} {n^a\xi_a} \\ \tilde{\Gamma}^{(T)}_i &=& 0 \\ V^{(L)}_{ij} &=& -\frac{9}{5n^a\xi_a} \left( e^{-4\phi}\xi_i\xi_j -\frac13\tilde{\gamma}_{ij} \right)\xi\!\cdot\!\tilde{A}\!\cdot\!\xi \\ V^{(T)}_{ijk} &=& -\frac{6}{5n^a\xi_a} \left( \xi_i\xi_j -\tilde{\gamma}_{k(i}\xi_{j)} \right)e^{-4\phi}\xi\!\cdot\!\tilde{A}\!\cdot\!\xi\\ \tilde{A}^{(LT)}_i &=& 0 \\ \tilde{A}^{(TT)}_{ij} &=& \frac{e^{-4\phi}}{2} (e^{-4\phi}\xi_i\xi_j-\tilde{\gamma}_{ij}) \xi\!\cdot\!\tilde{A}\!\cdot\!\xi \end{eqnarray} \end{mathletters} These are clearly four distinct eigenvectors, since there are four distinct values of $n^a\xi_a$ given by the roots of the determinant (\ref{chardet}). These can be thought as being labeled, essentially, by $K$ or $\xi\!\cdot\!\tilde{A}\!\cdot\!\xi$ indistinctly, or we can associate one characteristic speed to $K$ and the other one to $\xi\!\cdot\!\tilde{A}\!\cdot\!\xi$, as we prefer to do below. Summarizing, we have characteristic speeds obtained from the following distinct values of $n^a\xi_a$: \begin{description} \item[a)] $n^a\xi_a=0$, timelike, with eigenfields (essentially) $\tilde{\gamma}_{ij}, \phi, Q_i, \tilde{\Gamma}^{(T)}, V^{(T)}_{ijk}$. \item[b)] $n^a\xi_a=1$, null, with eigenfields (essentially) $\tilde{A}^{(TT)}_{ij}$. \item[c)] $n^a\xi_a=(1+\frac12(7c/10-3/5-2d(2a+1)))^{1/2}\equiv C_1$, with eigenfields (essentially) $\tilde{A}^{(TL)}_{i}$. \item[d)] $n^a\xi_a= (2a/3)^{1/2}\equiv C_2$, with eigenfield (essentially) $K$ \item[e)] $n^a\xi_a= (3/5-2bc/3+16bd/3-4d/3+7c/15)^{1/2}\equiv C_3$, with eigenfield (essentially) $\tilde{A}^{(LL)}$ \end{description} In the expressions for the characteristic speeds c),d) and e), the parameters $a,b,c,d$ are given in terms of $n_1, n_2, n_3$ via (\ref{parametric}). These speeds may be superluminal or causal, depending on the values of $n_1, n_2, n_3$. We can choose $n_1, n_2, n_3$ so that $C_1, C_2$ and $C_3$ are all equal to 1. This is achieved by setting \begin{mathletters}\label{nullens} \begin{eqnarray} n_1^2 &=& \frac{4}{15}\frac{280n_2^2+49+400n_2^4} {60n_2^2-49} \\ n_3^2 &=& \frac{6400 n_2^2}{60n_2^2-49} \end{eqnarray} \end{mathletters} for any value of $n_2^2$ greater than $49/60$. This means that there is a one-parameter family of well-posed conformally-decomposed systems with ``physical'' characteristics. From the analytical point of view, there does not appear to exist an argument for singling out a preferred value of $n_2$. It is likely that a preferred value of $n_2$ will be dictated by optimal numerical behavior. The expressions for $a,b,c,d$ in terms of $n_2$, with $n_1$ and $n_3$ as above (\ref{nullens}), are as follows: \begin{mathletters} \begin{eqnarray} a &=& \frac32, \\ b &=& -\frac{60 n_2^2-49}{4(7+20 n_2^2)}, \\ c &=& \frac{7}{10 n_2^2}, \\ d &=& -\frac{60n_2^2-49}{800n_2^2}. \end{eqnarray} \end{mathletters} \subsection{Propagation of the constraints} The propagation of the constraints can be calculated by taking a time derivative of each one of the constraint expressions, and subsequently using the evolution equations (\ref{combs}) to eliminate the time derivative of the fields in the right-hand side in favor of spatial derivatives, which recombine to yield back the constraints. We obtain \begin{mathletters}\label{subs1} \begin{eqnarray} \dot{\cal H} &=& \beta^l {\cal H},_l +(c-8d)\alpha e^{-4\phi}\tilde{\gamma}^{kl}{\cal P}_{k,l} + \cdots \\ \dot{\cal P}_i &=& \beta^l {\cal P}_{i,l} +\frac{\alpha}{6}(1-4b){\cal H},_i -\frac{\alpha}{2}e^{-4\phi} \left( \tilde{\gamma}^{jl}\left( \tilde{\gamma}^{kr}{\cal V}_{ijk,rl} -\frac{7}{10}\tilde{\gamma}_{im}{\cal G}^m,_{jl} \right) +\frac{1}{10}{\cal G}^m,_{mi} \right) \nonumber\\ && -2\alpha(2a+1)e^{-4\phi} \tilde{\gamma}^{jl} {\cal Q}_{[i,l]j} + \cdots \\ \dot{\cal G}^i &=& \beta^l {\cal G}^i,_l + \cdots \\ \dot{\cal Q}_i &=& \beta^l {\cal Q}_{i,l} + \cdots \\ \dot{\cal V}_{ijk} &=& \beta^l {\cal V}_{ijk,l} + \cdots \end{eqnarray} \end{mathletters} where $\cdots$ denotes undifferentiated terms proportional to the constraints themselves. To analyze the constraint propagation we proceed to turn (\ref{subs1}) into first order by defining several sets of variables which represent all the spatial derivatives of ${\cal V}_{ijk}, {\cal G}^i$ and ${\cal Q}_i$: \begin{mathletters} \begin{eqnarray} {\cal W}_{ij} &=& {\cal V}_{ijk},^k -\frac15{\cal G}_{(i,j)} +\frac{1}{15}\tilde{\gamma}_{ij}{\cal G}^k,_k \\ {\cal X}_{ijkl} &=& {\cal V}_{ijk,l} -\frac13\tilde{\gamma}_{kl}{\cal V}_{ijm},^m \\ {\cal U}_{ij} &=& {\cal Q}_{[i,j]} \\ {\cal Z}_{ij} &=& {\cal Q}_{(i,j)} \\ {\cal T}_{ij} &=& {\cal G}_{[i,j]} \\ {\cal J}_{ij} &=& {\cal G}_{(i,j)} +\frac{30}{7}A{\cal Q}_{(i,j)} \end{eqnarray} \end{mathletters} where $A$ is a constant which will be fixed shortly. Calculating the time derivative of these we obtain the resulting first-order system of evolution of the constraints: \begin{mathletters}\label{subs2} \begin{eqnarray} \dot{\cal H} &=& \beta^l {\cal H},_l +\alpha(c-8d) e^{-4\phi} {\cal P}_l,^l + \cdots \\ \dot{\cal P}_i &=& \beta^l {\cal P}_{i,l} +\frac{\alpha}{6}(1-4b){\cal H},_i -\frac{\alpha}{2}e^{-4\phi} {\cal W}_{il},^l -2\alpha(2a+1)e^{-4\phi} {\cal U}_{il},^l \nonumber \\ && -A\alpha e^{-4\phi} {\cal Z}_{il},^l +\frac{7\alpha}{30}e^{-4\phi} {\cal J}_{il},^l +\frac{11\alpha}{30}e^{-4\phi} {\cal T}_{il},^l + \cdots \\ \dot{\cal W}_{ij} &=& \beta^l {\cal W}_{ij,l} -\frac{c\alpha}{5} {\cal P}_{(i,j)} +\frac{c\alpha}{15} \tilde{\gamma}_{ij}{\cal P}_l,^l + \cdots \\ \dot{\cal X}_{ijkl} &=& \beta^m {\cal X}_{ijkl,m} + \cdots \\ \dot{\cal U}_{ij} &=& \beta^l {\cal U}_{ij} +\alpha d {\cal P}_{[i,j]} + \cdots \\ \dot{\cal Z}_{ij} &=& \beta^l {\cal Z}_{ij} +\alpha d {\cal P}_{(i,j)} + \cdots \\ \dot{\cal T}_{ij} &=& \beta^l {\cal T}_{ij} +\alpha c {\cal P}_{[i,j]} + \cdots \\ \dot{\cal J}_{ij} &=& \beta^l {\cal J}_{ij} +\alpha \left(c+\frac{30}{7}Ad\right) {\cal P}_{(i,j)} + \cdots \\ \dot{\cal G}^i &=& \beta^l {\cal G}^i,_l + \cdots \\ \dot{\cal Q}_i &=& \beta^l {\cal Q}_{i,l} + \cdots \\ \dot{\cal V}_{ijk} &=& \beta^l {\cal V}_{ijk,l} + \cdots \end{eqnarray} \end{mathletters} For this system there is a symmetrizer given by \begin{eqnarray}\label{constrenergy} \bar{u} \mbox{\boldmath$H$}_C u &=& e^{4\phi}\frac{(1-4b)}{6(c-8d)} {\cal H}^2 + {\cal P}_i{\cal P}^i +e^{-4\phi}\frac{5}{2c} {\cal W}_{ij}{\cal W}^{ij} + {\cal X}_{ijkl}{\cal X}^{ijkl} -e^{-4\phi}\frac{2(2a+1)}{d}{\cal U}_{ij}{\cal U}^{ij} \nonumber\\ && -e^{-4\phi}\frac{A}{d} {\cal Z}_{ij}{\cal Z}^{ij} +e^{-4\phi}\frac{11}{30c} {\cal T}_{ij}{\cal T}^{ij} +e^{-4\phi}\frac{7/30}{c+Ad30/7} {\cal J}_{ij}{\cal J}^{ij} \nonumber\\ && + {\cal G}_i{\cal G}^i + {\cal Q}_i{\cal Q}^i + {\cal V}_{ijk}{\cal V}^{ijk}. \end{eqnarray} Taking $A=\frac{7}{60}(-c/d)$, $\mbox{\boldmath$H$}_C$ is positive definite because, under the conditions (\ref{parametric}), all the factors accompanying the squares of the fields are strictly positive. This shows that no additional restrictions on the ranges of the parameters $a,b,c,d$ are necessary in order to have well posed constraint evolution. \section{Conclusion }\label{sec:4} We have derived a 3-parameter family of well posed versions of the conformally-decomposed 3+1 equations, perhaps amenable to successful numerical integration. One might object that there is no need for it in view of the results in~\cite{BS99}, but we can argue rather strongly that these results may prove helpful in pinning-down the main cause of numerical instabilities. This well posed version requires the lapse to be proportional to the determinant of the intrinsic geometry of the surfaces, and requires combinations of the contraints with the evolution equations. The lapse density $\sigma$ and the shift vector $\beta^i$ are arbitrary non-dynamical variables, which means that they must be specified as free source functions. This well posed version uses the same variables as the original system (except for the addition of the first spatial derivatives of the densitized 3-metric, referred to as ``conformal metric'' by the authors in~\cite{BS99}). In addition, this well-posed version of the original equations propagates the constraints in a stable manner, which is relevant to unconstrained evolution. We think that this is the least invasive way to turn the original conformally-decomposed system into a well posed one. In practice, a choice of the numerical parameters $n_1, n_2, n_3$ must be made. The characteristic speeds depend on this choice. Optimal choices of the parameters $n_1, n_2, n_3$ for numerical evolution are those that ensure that the characteristics are all either null or timelike. With such a choice, the formulation would be suited to evolve blackhole spacetimes outside the event horizon. Among these choices, it has been suggested~\cite{fixing} that the preferred one would be the one for which the characteristics are all ``physical'', namely, either null or normal to the slices. We have shown that such a choice is possible for an arbitrary $n_2>\sqrt{49/60}$. The systems obtained in this work are not contained in our previous work~\cite{frittelliletter,frittelli-reula99}. The choice of variables in~\cite{frittelliletter,frittelli-reula99} is inadequate for decomposing the trace and trace free part of the extrinsic curvature, as well as for extracting the determinant of the 3-metric. This is clear from the fact that, in that work, the available parameters $\alpha$ and $\beta$ are not allowed to take the value $-1/3$ without the argument breaking down. The systems obtained here differ significantly from the system obtained in~\cite{brugman} by considering the trace of the extrisic curvature $K$ and the determinant of the instrinsic metric (and its derivatives) as dynamical variables on equal footing with the rest, rather than as free source functions. Furthermore, we have obtained a 3-parameter family of systems, one system for each appropriate choice of $n_1, n_2, n_3$, whereas in~\cite{brugman} there is only one system which preserves the trace conditions. Additionally, we have separated the divergence of the intrinsic metric $\tilde{\Gamma}^i$ from the divergence-free part of the metric. This decomposition keeps up with the spirit of~\cite{BS99}. We have found that in obtaining these well-posed formulations the lapse must be proportional to some power of the determinant of the intrinsic metric, since the parameter $a$ cannot take the value 0. This is similar to our findings in~\cite{frittelli-reula94,frittelliletter,frittelli-reula99}, as well as other notable cases~\cite{helmut96,mirta97,mirta98,york98}. In our present case this is remarkable, since we have used quite general energy norms. \acknowledgments We thank Roberto G\'{o}mez for enlightening discussions. This work has been supported by NSF under grant No. PHY-9803301, and by CONICOR, CONICET, and SeCyT, UNC.
\section{Introduction} \vspace*{-0.5pt} \noindent The scalar mesons appearing around 1 $GeV$ mass scale seems to be least understood among spin zero mesons. The experimental data\cite{1}, on strong decays of $a_{0}(980)$, $f_{0}(980)$ and $f_{0}(1370)$ do not lead to the final understanding of the real structure of these mesons \cite{2}$-$\cite{PLB98}. Ideas exists that these states are $K \bar K$ molecules \cite{2}$-$\cite{7}. There is a suggestion that $a_0(980)$ has a four-quark structure with the strange quarks ($s \bar s$) contribution\cite{Achasov}, based on the analyses of the $\phi \to \gamma a_0(980) \to \gamma \eta \pi$\cite{PLB98}. Such uncertainties suggest further investigation, which were concentrated on $2 \gamma$ decays of $a_{0} (980)$, $f_{0}(980)$, $f_{0}(1370)$ and $\chi_{c0}$ mesons. The aim was to explain the experimental decay widths and at the same time to reproduce the measured masses. The relativistic quark model is used to correlate various data and to establish the connection between masses and decay widths. For that purpose one employs the covariant model \cite{16,18} which includes the heavy - quark symmetry. As shown in the following section the model and the calculations are both covariant and gauge invariant. This model is a covariant generalization \cite{16,18} of the well known ISGW model \cite{17}. However, here the usage of small quark masses is investigated, which means the avoidance of the weak binding limit approximation in its strictest sense \cite{17}. That might better mimick the real quark fields which should appear in the photon emitting quark loop (Fig. 1) in the first order of QED/QCD expansion. In a very simplified version of that model, which is employed here, only the quark momentum distribution parameter $\beta$ and the model quark masses appear. All parameters are correlated and compared with the nonrelativistic choices\cite{17,19,20} The quark masses,were treated as fitting parameters in a limited sense. Suitable values, allowed within the experimental uncertainty in current quark masses \cite{1}, were selected. That parameterization is expected to lead to a reasonable reproduction of meson masses. No additional fitting was allowed when widths were calculated. In that way one can reproduce the measured $\Gamma (2 \gamma)$ reasonably well and make the prediction for the $f_{0}(1370)$ $2 \gamma$ decay. All results are based on the valent quark $q\overline{q}$ structure, which is characteristic of the model. The importance of $q\overline{q}$ structure has often been mentioned \cite{7}$-$\cite{14}. Our basic loop diagrams, Fig. 1 below, correspond closely to the quark loop diagrams shown in Fig. 1 of Ref.(10). Thus it is not surprising that our results, following from the somewhat more complicated diagrams, depends strongly on quark masses. Moreover our model contains the sea component constrained by the requirement of general Lorentz covariance, valid in an frame \cite{16,18}. In that way, indirectly, some other QCD structures, as discussed earlier \cite{16,18}, enter into our description. Predictions based on our simplified model version test how well, or how badly the model mimicks the real QED/QCD world. The quarkonium approximation \cite{17} is investigated here in the circumstances which are different from the usual form factors related problems\cite{16,17}. The results depend also on the quark flavor structure of the scalar mesons. They do distinguish among various propose $u\overline{u}, d\overline{d}$ and $s\overline{s}$ mixing\cite{13,14} in $f_{0}$ mesons.We investigate only lower laying states without entering into discussion of the states as $a_0(1450)$ which would require to include the higher order terms of our model. \newpage \vspace*{1pt} \section{Brief description of the model} \vspace*{-0.5pt} \setcounter{section}{2} \renewcommand{\thesection.\arabic{equation}}{\arabic{section}.\arabic{equation}} \noindent A scalar meson $H$ with the four-momentum $P$ and the mass $M$ is covariantly represented \cite{16,18} by \begin{displaymath} |H(E,\vec{P}, M) \big> = \frac {N(\vec{P})}{(2\pi)^{3}} \sum_{c,s_{1},s_{2}} \ \ \sum_{f} C_{f} \ \ \int \ \big[ 4m_{1}m_{2} \big] \ d^{4}p \ \delta (p^{2}-m^{2}_{1}) \ \Theta (e) \end{displaymath} \begin{equation} \cdot d^{4}q \ \delta(q^{2}-m^{2}_{2}) \ \Theta (\epsilon) \ d^{4}K \ F(K)\ \delta^{(4)} (p+q+K-P) \ \Theta (E) \ \phi_{f} (l_{\perp}) \end{equation} \begin{displaymath} \cdot \overline{u}_{f,s_{1}}^{c} (\vec{p}) v^{c}_{f,s_{2}} (\vec{q}) \ \ d^{+}_{f} (\vec{q}, c, s_{2}) b^{+}_{f} (\vec{p}, c, s_{1}) | 0 \big> \end{displaymath} Here $m_{i}$ are quark masses and $f$ stands for quark flavor. The quark wave function is \begin{equation} \phi_{f}(l^{\mu}_{\bot}) = \frac{1} {(1 - \frac{l_{\bot}^{2}}{4\beta_{f,H}^{2}})^{2}} \end{equation} \begin{displaymath} l^{\mu}_{\bot} (P) = l^{\mu} - \frac{P^{\mu}(P \cdot l)}{M^{2}} \end{displaymath} with $l^{\mu} = (p -q)^{\mu}/2$, $p^{\mu}=(e,\vec{p})$ and $q^{\mu}=(\epsilon,\vec{q})$. The fitting parameter $\beta_{f,H} \ \ (f=u,d,s,c)$ is fixed by fitting the meson mass as described below. The dipole form (2.2) was found to be a better choice than the exponential form used earlier \cite{16,18}. (See also some remarks in Appendix.) Coefficient $C_{f}$ indicates the flavor content of a particular meson (For example, see below formula (3.14), where for $f_{0}(980) \ \ C_{u}=cos(\theta)/ \sqrt{2}$). The symbols $\overline{u}, v, d^{+}, b^{+}$ correspond to valence quarks while the sea function has a general form \begin{equation} F(K) = \delta^{(4)} \big[K^{\mu} - \frac{P^{\mu}}{M} \frac{P^{\nu}\big(P- p - q\big)_{\nu}}{M} \big] \ \varphi(K). \end{equation} For simplicity we set \begin{equation} \varphi(K)=1. \end{equation} The complex looking Dirac function in (2.3) simplifies the model structure easing all formal manipulations. One could produce a somewhat more complicated model, without that Dirac function. Additional parameter(s) in the sea model function would lead to a richer and more flexible model \cite{18}. Thus the choice (2.4) correspond to a minimalistic model version. The scalar meson state (2.1) is normalized so that the matrix element of the vector current $V^{\mu}$ would be, for example, \begin{equation} \langle H (\vec{P}_{f})| V^{\mu} | H (\vec{P}_{i})\rangle = \frac{1}{(2 \pi)^{3}} F_{+}(Q^{2}) (P_{f} +P_{i})^{\mu} + \cdots \end{equation} with $Q^{\mu}=(P_{f} - P_{i})^{\mu}$, $F_{+}(Q^{2}=0) =1$. The normalization (2.5) insures that the vector current, and thus charge, is conserved. This requirement is equivalent to the condition \begin{displaymath} \langle H (E,\vec{P},M)| H (E,\vec{P},M)\rangle = 2E = \end{displaymath} \begin{displaymath} =\frac {N(\vec{P})^{2}}{(2\pi)^{6}} \sum_{c,s_{1},s_{2}} \sum_{ c',s'_{1},s'_{2}} \sum_{f,f'} C_{f} C_{f'} \cdot \int d^{3}p \frac{m_{1}m_{2}}{e} \frac{M}{E} \frac{\phi_{f} (l_{\bot})}{q_{\parallel}} d^{3}p' \frac{m_{1}m_{2}}{e'} \frac{M}{E} \frac{\phi_{f'}(l'_{\bot})}{q'_{\parallel}} \end{displaymath} $$ \cdot[ \overline{ v}_{f',c',s'_{2}}(\vec{q'}) u_{f',c',s'_{1}} (\vec{p'}) \overline{u}_{f,c,s_{1}} (\vec{p}) v_{f,c,s_{2}} (\vec{q})]\cdot $$ \begin{equation} \cdot \big<0| b_{f'} (\vec{p'}, c', s'_{1}) d_{f'} (\vec{q'}, c', s'_{2}) d^{+}_{f} (\vec{q}, c, s_{2}) b^{+}_{f}(\vec{p}, c, s_{1}) | 0 \big> \big|_{ \vec{q}= T_1 , \ \ \vec{q'}= T_1' } \end{equation} Here $$T_1=- \vec{p} + \frac{\vec{P}}{M} (p_{\parallel})T$$ \begin{equation} T=1+ \frac{ \sqrt{m^{2}_{2}-m^{2}_{1}+p^{2}_{\parallel}} } {p_{\parallel}} \ \ \ ;\ \ \ \ \ p_{\parallel}= \frac{P \cdot p}{M} \ \ \ ;\ \ \ \ \ q_{\parallel}= \frac{P \cdot q}{M} \end{equation} After a lengthy but straightforward manipulation, one find \begin{displaymath} N(\vec{P}) = \frac{E}{M} N(0) \end{displaymath} \begin{equation} \langle H (E,\vec{P},M)| H (E,\vec{P},M)\rangle = 2E =3N(0)^{2} \sum_{f} C_{f}^{2} \int d^{3}p \ \frac{\epsilon}{e} \ {\big(}\frac{\phi_{f} (l_{\bot})}{q_{\parallel}}{\big)}^{2} \ (pq -m_{1}m_{2}) \end{equation} >From (2.8) $N(0)$ can be calculated numerically. The matrix element of the conserved vector current $V^{\mu}$ has to vanish when current acts on the scalar meson state, i.e. \begin{equation} \langle 0| V^{\mu} | H (E, \vec{P},M) \rangle \equiv 0 \end{equation} The model states (2.1) are consistent with this very general requirement. Some additional details are shown in the Appendix. So far the model is closely related to ISGW model \cite{17}. In the nonrelativistic limit and in the weak binding approximation it goes exactly in the ISGW form\cite{16,18}. However the weak binding approximation means that the quark masses and the quark energies are approximately equal \cite{17}. In the present application, the model quark fields enter a loop (Fig. 1) which constitutes the lowest QED approximation. The QCD corrections are modeled by the functions (2.2) and (2.3). One can try to mimick the real QED/QCD world by retaining small (current) quark masses in the model. Then the meson mass should be equal to a sum (weighted by the sea influence) of average model quark energies. In the model determined by (2.2) and (2.4) this i \begin{equation} M= \frac {3 N(0)^{2}}{2E} \sum_{f} C_{f}^{2} \int d^{3}p \ (\frac{\phi_{f} (l_{\bot})}{q_{\parallel}})^{2} \ \frac{p \cdot q - m_{1} m_{2}}{e/ \epsilon} (p_{\parallel} + q_{\parallel})|_{ \vec{q}= - \vec{p} + \frac{\vec{P}}{M} (p_{\parallel})T} \end{equation} As discussed in the Appendix this simple form holds in the minimalistic model version (2.4). The wave function $\phi_p$ can be connected with the usual potential$^{17,19,20}$ as shown in Appendix. In the nonrelativistic, weak binding limit (WBL) (2.10) goes into \begin{equation} M \cong \langle (e+\epsilon) \rangle \stackrel{WBL}{\rightarrow }\hat{m}_{1}+\hat{m}_{2} \end{equation} Here $\hat{m}_{i}$ are constituent quark masses \cite{17}. This WBL makes sense only if one uses constituent masses $\hat{m}_{i}$, with the corresponding $\beta $'s \cite{17} in all relevant formulae. If (2.10) is calculated explicitly in our model it can hold only for particular values of model parameters, i.e. $\beta_{f,H}$ with a particular set of quark masses. When one aims for $m_{i}$ close to the current quark masses, that requires the consistent $\beta_{f,H}$ values. As explained in Appendix, by using Eq.(2.10), one determines the theoretical form factor at the physical momentum transfer $Q^2$. However, one is still dealing with some sort of a mock meson description. Small quark masses, which are used with the relativistic model, lead to the $\beta$ values which are close to those used in the nonrelativistic model. Full comparisment between those cases is given in Appendix. \newpage \section{Electromagnetic widths} \setcounter{equation}{0} \setcounter{section}{3} \renewcommand{\thesection.\arabic{equation}}{\arabic{section}.\arabic{equation}} \noindent In our model \cite{16,18} the amplitude ${\cal M}$ for the transition $f_{0} \rightarrow 2\gamma$ is determined from the leading diagrams shown in Fig. 1. \begin{figure}[ht] \begin{center} \includegraphics[width=8.5cm,height=4.78cm]{fig1acad.eps} \caption{Two-photon decay. Full lines are valence quarks, wavy lines are emitted photons and blobs symbolize scalar meson states.} \end{center} \end{figure} Although the diagrams in Fig. 1 closely resemble the free quark diagrams, they are not the same. One has to sum over the moment $\vec{p}$, $\vec{q}$ and over the spins of the valence quark states. These sums are weighted by the corresponding functions in the meson state (2.1). A loop corresponds to each quark flavor. For example, for the flavor $d$ the amplitude corresponding to the diagram in Fig. 1a is determined by \newpage \begin{displaymath} M_{1}^{\mu \nu} = (2\pi)^{3/2} \langle 0|:\overline {\Psi_{d}} \gamma^{\mu} S_{F}(l_{2} - k_{1})\gamma^{\nu} \Psi_{d}:|(d \overline {d})\rangle = \end{displaymath} \begin{displaymath} = (2\pi)^{3/2} \sum_{c_{1},\alpha,\beta} \langle 0|: \int \frac {d^{3}l_{1}}{(2\pi)^{3}} \frac {m_{d}}{l_{f}^{0}} \big[ b_{d}^{+}(\vec{l_{1}},\alpha,c_{1}) \overline {u_{d}}(\vec{l_{1}},\alpha) + d_{d}(\vec{l_{1}},\alpha,c_{1})\overline {v_{d}}(\vec{l_{1}},\alpha) \big] \end{displaymath} \begin{equation} \cdot \gamma^{\mu} \frac{\not \! l_{2} - \not \! k_{1} +m_{d}}{(\not \! l_{2} - \not \! k_{1})^{2} - m_{d}^{2}} \gamma^{\nu} \end{equation} \begin{displaymath} \cdot \int \frac {d^{3}l_{2}}{(2\pi)^{3}} \frac {m_{d}}{l_{2}^{0}} \big[ b_{d}(\vec{l_{2}},\beta,c_{1})u_{d}(\vec{l_{2}},\beta) + d^{+}_{d}(\vec{l_{2}},\beta,c_{1})v_{d}(\vec{l_{2}},\beta) \big]: \end{displaymath} \begin{displaymath} \cdot \frac {N(\vec{P})}{(2\pi)^{3}} \sum_{c, s_{1},s_{2}} \int d^{3}p \frac{m_{d}^{2}}{e} \frac{M}{E} \frac{\phi_{d, f_{0}} (l_{\bot})}{q_{\parallel}} \overline{u}_{d,c,s_{1}} (\vec{p}) v_{d,c,s_{2}} (\vec{q}) d^{+}_{d} (\vec{q}, c, s_{2}) b^{+}_{d}(\vec{p}, c, s_{1}) | 0 \big> \big|_{ \vec{q}= T_1 } \end{displaymath} The contraction of the creation (annihilation) operators in (3.1) leads to the summation over spin indicies. That give \begin{displaymath} M_{1}^{\mu\nu} = -3 \frac{N(0)}{(2\pi)^{3/2}} \int d^{3}p \frac{m_{d}^{2}}{e} \frac{\phi_{d, f_{0}} (l_{\bot})}{q_{\parallel}} \end{displaymath} \begin{equation} \cdot Tr [\gamma^{\mu} \frac{\not \! p - \not \! k_{1} +m_{d}}{(-2 p k_{1})} \gamma^{\nu} \frac{\not \! p +m_{d}}{2 m_{d}} \frac{\not \! q - m_{d}}{2 m_{d}}] \big|_{ \vec{q}= - \vec{p} + \frac{\vec{P}}{M} (p_{\parallel})T } \equiv I(\vec{p},\vec{q}) Tr(k_{1})^{\mu\nu} \end{equation} The amplitude corresponding to Fig. 1 is \begin{equation} {\cal M}= M_{1}^{\mu\nu} \epsilon_{\mu}(k_{2},\lambda_{2})\epsilon_{\nu}(k_{1},\lambda_{1}) + M_{2}^{\mu\nu} \epsilon_{\mu}(k_{1},\lambda_{1})\epsilon_{\nu}(k_{2},\lambda_{2}), \\ \end{equation} \begin{displaymath} M_{2}^{\mu\nu}=I(\vec{p},\vec{q}) Tr(k_{1} \rightarrow k_{2})^{\mu\nu} \end{displaymath} Routine calculation of traces produces the final result which contains \begin{equation} Tr(k_{1})^{\mu\nu}=A_{1}(2p^{\nu}q^{\mu} - 2 p^{\mu} p^{\nu} + k_{1}^{\mu} (p-q)^{\nu} + k_{1}^{\nu} (p-q)_{\mu} - g^{\mu\nu} \big[k_{1} (p-q)\big]) \end{equation} \begin{displaymath} A_{1,2}= \frac{1}{2m_{d}(p \cdot k_{1,2})} \end{displaymath} It is convinient to carry out further calculation in the meson rest frame (MRF). That is defined by \begin{displaymath} P^{\mu}= (M, \vec{0}) \ ; \ \ \ e=\epsilon \ ; \ \ \ \vec{q}= - \vec{p} \end{displaymath} \begin{equation} k_{1}^{\mu} = (\omega, \vec{k}) \ ; \ \ \ k_{2}^{\mu} = (\omega, -\vec{k}) \ ; \ \ \ \omega = |\vec{k}| = \frac {M}{2} \end{equation} Further simplification is obtained by selecting orthogonal polarization vectors \begin{displaymath} P \cdot \epsilon_{1,2} = k_{1} \cdot \epsilon_{1} = k_{2} \cdot \epsilon_{2} = 0 \end{displaymath} \begin{equation} \epsilon^{\mu}(k_{2},\lambda_{2}) \equiv \epsilon^{\mu}_{2} = (0, \vec{\epsilon}_{2}) \ ; \ \ \ \epsilon^{\mu}(k_{1},\lambda_{1}) \equiv \epsilon^{\mu}_{1} = (0, \vec{\epsilon}_{1}) \ ; \ \ \ k_{2} \cdot \epsilon_{1} = k_{1} \cdot \epsilon_{2} = 0 \end{equation} As shown in the next section the result does not depend on a particular gauge. With (3.5), (3.6) one obtain \begin{equation} {\cal M}= -2I(\vec{p},\vec{q})\{ A_{1} \big[(\vec{\epsilon}_{1}\vec{\epsilon}_{2}) (\vec{k} \vec{p}) + 2(\vec{p}\vec{\epsilon}_{1}) (\vec{p}\vec{\epsilon}_{2})\big] + A_{2}\big[-(\vec{\epsilon}_{1}\vec{\epsilon}_{2}) (\vec{k} \vec{p}) + 2(\vec{p}\vec{\epsilon}_{1}) (\vec{p}\vec{\epsilon}_{2})\big] \} \end{equation} \begin{displaymath} I(\vec{p},\vec{q})|_{\vec{P}=0} = -3(2\pi)^{3/2} N'(0) \int d^{3}p \frac{m_{d}^{2}}{e^{2}} \frac{1}{(1 + \frac{\vec{p}^{2}}{4\beta_{d,f_{0}}^{2}})^{2}} \end{displaymath} By using $\vec{p}\vec{k} = p \omega cos\theta$ one obtains \begin{displaymath} {\cal M}= - \frac{3 N(0)}{m_{d}\omega (2\pi)^{1/2}} \int p^{2}dp \ sin\theta d \theta \ \frac{m_{d}^{2}}{e^{2}} \frac{1}{(1 + p^{2}/(4\beta_{d,f_{0}}^{2}))^{2}} (\vec{\epsilon}_{1}\vec{\epsilon}_{2}) \frac{2 p^{2} (\omega cos^{2}\theta + e sin^{2}\theta)}{e^{2} - p^{2} cos^{2}\theta } \end{displaymath} \begin{equation} \equiv (\vec{\epsilon}_{1}\vec{\epsilon}_{2}) \cdot I_{d,\overline{d}}(f_{0}) \end{equation} The calculation of the decay width \begin{equation} \Gamma = \frac{1}{32 \pi M_{H}} \overline{ |{\cal M}_{H}|^{2} } \end{equation} requires the summation over photon polarization states, \begin{equation} \overline{\sum_{\lambda_{1}\lambda_{2}} |{\cal M}|^{2} } = 2 \cdot I_{f,\overline{f}}(H)^{2} \end{equation} as well as the summation over quark flavors in (3.8), connected with the meson quark structure, which we parameterize a $|H\rangle = \sum_{f} C_{f} |f,\overline{f}\rangle$. \begin{displaymath} |f_{0}(980)\rangle = \frac{cos\theta}{\sqrt{2}}(|u \overline{u}\rangle + |d \overline{d}\rangle) + sin\theta |s \overline{s}\rangle \end{displaymath} \begin{displaymath} |a_{0}(980)\rangle = \frac{1}{\sqrt{2}}(|u \overline{u}\rangle - |d \overline{d}\rangle) \end{displaymath} \begin{displaymath} |f_{0}(1370)\rangle = \frac{-sin\theta}{\sqrt{2}}(|u \overline{u}\rangle + |d \overline{d}\rangle) + cos\theta |s \overline{s}\rangle \end{displaymath} \begin{equation} |\chi_{c_{0}}(3415)\rangle = |c \overline{c}\rangle \end{equation} The physical mixing for these states is usually determined \cite{14} by \begin{equation} cos\theta = 1 \ ; \ \ \ sin\theta = 0 \end{equation} or \begin{equation} cos\theta = \frac{1}{3} \ ; \ \ \ sin\theta = \frac{2\sqrt{2}}{3} \end{equation} Eventually one finds by summing over flavors: \begin{displaymath} Int(f_{0}(980)) =\frac{e^{2}}{9} [\frac{5cos\theta}{\sqrt{2}} I_{u \overline{u}}(f_{0}(980)) + sin\theta I_{s \overline{s}}(f_{0}(980))] \end{displaymath} \begin{displaymath} Int(a_{0}(980)) =\frac{e^{2}}{3\sqrt{2}} I_{u \overline{u}}(a_{0}(980)) \end{displaymath} \begin{displaymath} Int(f_{0}(1370)) =\frac{e^{2}}{9} [\frac{-5sin\theta}{\sqrt{2}} I_{u \overline{u}}(f_{0}(1370)) + cos\theta I_{s \overline{s}}(f_{0}(1370))] \end{displaymath} \begin{equation} Int(\chi_{c_{0}}(3415)) = \frac{4e^{2}}{9}I_{c \overline{c}}(\chi_{c_{0}}(3415)) \end{equation} where $I_{f,\overline{f}}(H)$ is given by formula (3.8) and $e$ is the electron unit charge , i.e. $e^{2}/(4\pi) = \alpha = (137,04)^{-1}$. The $2\gamma$-decay width can be put in the final form valid for a meson $H$ from (3.11): \begin{equation} \Gamma(H) = \frac{\pi \alpha^{2}}{M_{H}} (\frac{Int(H)}{e^{2}})^{2} \end{equation} \newpage \section{Gauge invariance explicitly tested} \setcounter{equation}{0} \setcounter{section}{4} \renewcommand{\thesection.\arabic{equation}}{\arabic{section}.\arabic{equation}} \noindent Through its covariant nature, being in a sense a simplified rendering of the real QCD field theory, the model \cite{16,18} automatically produces gauge invariant results. Under the gauge transformation \begin{equation} \epsilon_{i \mu} \rightarrow \epsilon_{i \mu} + \Lambda k_{i \mu} \end{equation} the amplitude (3.3) should not change. That leads to the equality \begin{displaymath} Tr(k_{1})^{\mu\nu} (\epsilon_{2} + \Lambda k_{2})_{\mu} (\epsilon_{1} + \Lambda k_{1})_{\nu} + Tr(k_{1} \rightarrow k_{2})^{\mu\nu} (\epsilon_{1} + \Lambda k_{1})_{\mu} (\epsilon_{2} + \Lambda k_{2})_{\nu} = \end{displaymath} \begin{equation} = Tr(k_{1})^{\mu\nu} \epsilon_{2 \mu} \epsilon_{1 \nu} + Tr(k_{1} \rightarrow k_{2})^{\mu\nu} \epsilon_{1 \mu} \epsilon_{2 \nu} + N(\Lambda, k_{1}, k_{2}) \end{equation} Here first two terms give the amplitude (3.3). The piece $N(\Lambda, k_{1}, k_{2})$ should not contribute to the physical amplitude (3.7). In the meson rest frame, using (3.5) and (3.6), one immediately finds \begin{equation} N(\Lambda, k_{1}, k_{2}) = \Lambda \frac{2\omega}{m_{d}} (\frac{\vec{p} \vec{\epsilon_{1}}}{p \cdot k_{1}} + \frac{\vec{p} \vec{\epsilon_{2}}}{p \cdot k_{2}}) [\omega - e(\vec{p})] \end{equation} while the terms proportional to $\Lambda^{2}$ cancel. $N(\Lambda, k_{1}, k_{2})$ enters the integration over the bound quark momentum $\vec{p}$, as shown in (3.2), (3.3). The corresponding change in ${\cal M}$ is \begin{equation} \Delta{\cal M} = I(\vec{p}, \vec{q})|_{\vec{P}=0} \cdot N(\Lambda, k_{1}, k_{2}). \end{equation} Here one has \begin{displaymath} \int d^{3}p = \int_{0}^{\infty} p^{2}dp \int_{0}^{\pi} sin\theta d\theta \int_{0}^{2\pi} d\Phi \end{displaymath} \begin{equation} \vec{p} \vec{\epsilon_{1,2}} =|\vec{p}| [ sin\theta cos\Phi (\hat{x}\vec{\epsilon_{1,2}}) + sin\theta sin\Phi (\hat{y}\vec{\epsilon_{1,2}}) ] \end{equation} Integration over the azimuthal angle $\Phi$ gives zero result, so one has \begin{equation} \Delta{\cal M} \equiv 0 \end{equation} as required by the gauge invariance. \newpage \section{Results and discussion} \setcounter{equation}{0} \setcounter{section}{5} \renewcommand{\thesection.\arabic{equation}}{\arabic{section}.\arabic{equation}} \noindent The application of the model (2.1) starts with the self-consistency condition (SCC) (2.10). Quark masses are selected, as close as practical to the Particle Data \cite{1} values. Than parameters $\beta_{f,H}$ are varied for various flavors appearing in (3.11) so as to reproduce the experimental meson masses \cite{1}. \begin{displaymath} M[f_{0}(980)] = 0.980 \ GeV \ ;\ \ \ \ \ \ M[a_{0}(980)] = 0.980 \ GeV \end{displaymath} \begin{equation} M[f_{0}(1370)] = 1.370\ GeV \ ;\ \ \ \ \ \ M[\chi_{c_{0}}(3415)] = 3.415 \ GeV \end{equation} Theoretical expression (2.10), (2.11) is a sum of parts corresponding to various flavors. For example \begin{equation} M[a_{0}(980)]_{th} = \frac{1}{2} (M_{u} +M_{d}). \end{equation} It turns out that SCC requires light quark masses somewhat larger than Particle Data \cite{1} median values, while strange and charm masses could be kept within Particle Data limit. The satisfactory selection i \begin{equation} m_{u} = m_{d} = 0.015 \ GeV \;\ \ \ m_{s} = 0.120 \ GeV \;\ \ \ m_{c} = 1.5 \ GeV \end{equation} The most stringent restriction on the light quark parameters is obtained by fitting the mass of the $a_{0}$ meson. If there is no strangeness mixing, i.e. with $\theta = 0$ in (3.11), the same mass is calculated for the $f_{0}(980)$ meson too. Some interesting values are shown in Table 1. \begin{table}[h] \caption{Mock mass values for $a_{0}$} \begin{center} \begin{tabular}{|c|c|} \hline $\beta_{u,d} (GeV)$ & $M_{0} (GeV)$ \\ \hline 0.260 & 0.885 \\ 0.270 & 0.919 \\ 0.280 & 0.953 \\ 0.288 & 0.980 \\ 0.295 & 1.004 \\ \hline \end{tabular} \end{center} \end{table} With ideal mixing $(\theta = 0)$ $f_{0}(1370)$ has a pure $s \overline{s}$ configuration. The corresponding mass values are shown in Table 2. \begin{table}[h] \caption{Mock mass values for $s \overline{s}$ configuration} \begin{center} \begin{tabular}{|c|c|} \hline $\beta_{s} (GeV)$ & $M_{0} (GeV)$ \\ \hline 0.350 & 1.272 \\ 0.381 & 1.370 \\ 0.400 & 1.435 \\ 0.450 & 1.599 \\ 0.500 & 1.764 \\ \hline \end{tabular} \end{center} \end{table} The mass of the pure $c \overline{c}$ state $\chi_{c_{0}}$ can be reproduced by using $\beta_{c} = 0.267 \ GeV$ as shown in Table 3. \begin{table}[h] \caption{Mock mass values for $c \overline{c}$ configuration} \begin{center} \begin{tabular}{|c|c|} \hline $\beta_{c} (GeV)$ & $M_{0} (GeV)$ \\ \hline 0.250 & 3.377 \\ 0.267 & 3.415 \\ 0.280 & 3.445 \\ 0.300 & 3.490 \\ 0.330 & 3.560 \\ \hline \end{tabular} \end{center} \end{table} When non\-ideal mixing (3.11), (3.13) is allowed the masses of $f_{0}(980)$ and $f_{0}(1370)$ can be reproduced by $\beta_{u}$ and $\beta_{s}$ which are different from those shown in Tables 1 and 2. However, the values in Table 1 still correspond to the $a_{0}$ mass. The masses $M[f_{0}(980)]$ and $M[f_{0}(1370)]$ are reproduced by: \begin{equation} \beta_{u,d} = 0.419 \ GeV \ ;\ \ \ \ \ \ \beta_{s} = 0.242 \ GeV \end{equation} The corresponding $\Gamma(H \rightarrow 2\gamma)$ values are summarized in Table 4. \begin{table}[h] \caption{Decay widths} \begin{center} \begin{tabular}{|c|c|c|c|} \hline Meson & Mixing & $\Gamma_{theory} (keV)$ & $\Gamma_{exp} (keV)$ \\ \hline $a_{0}(980)$ & & 0.137 & $0.26 \pm 0.08 $ \\ $f_{0}(980)$ & (3.12) & 0.380 & $0.56 \pm 0.11$ \\ $f_{0}(980)$ & (3.13) & 0.534 & $0.56 \pm 0.11$ \\ $f_{0}(1370)$ & (3.12) & 0.348 & \\ $f_{0}(1370)$ & (3.13) & 0.145 & \\ $\chi_{c_{0}}(3415)$ & & 4.608 & $4.0 \pm 2.8$ \\ \hline \end{tabular} \end{center} \end{table} All conclusions depend strongly on the quark masses. For example, if one chooses $m_{c} = 1.4\ GeV$ than the SCC requires $\beta_{c} = 0.346 \ GeV$. The $q \bar q$ structure, which is the main feature of the model, might be capable of explaining two photon decay. By that one does not mean a naive "free-quark" structure of an early nonrelativistic model. In the present model the valence quarks are immersed in a sea, which, however rudimentary, takes into the account the interference of other QCD induced configurations like for example $s\bar{s}$ pairs, gluons etc. The SCC (2.10) transmits that into the numerical results. The experimental error in $f_{0}(980) \to 2 \gamma$ rate is rather large. Although, the large theoretical prediction in Table 4, seems to be in better agreement with experiments, the smaller one, which corresponds to the ideal mixing, cannot be ruled out. However, the $f_0(1370)$ decay into pions indicates the presence of the light $q \bar q$ combinations\cite{7,10,11}. Our result also agrees with the nonideal mixing as considered by Lanik\cite{13}. If the corresponding theoretical predictions (Table 4) for the decay of $f_{0} (1370)$ turns out to be at least approximately correct, one would have a very strong support for nonideal mixing\cite{13,14}. The experimental data \cite{1} for the $a_{0}(980) \to 2 \gamma$ decay width contain large errors. Our theoretical value, Table 4, is close to the lower experimental limit. Various other theoretical approaches are summarized in Ref.(9). Our approach has some analogy with Deakin et al.\cite{10} who used constituent quark masses and concluded that theoretical results depend strongly on the numerical values of those masses. The same strong dependence on the masses, was found here. The decay width and the mass of $\chi_{c_{0}}$ are very well reproduced within the model. Naturally all our conclusions depend on the validity of model as such. Here we have tried a simplicistic version of the model, which relies on the functions (2.2) and (2.4) That gave SCC (2.10) which represents a strong restriction on the model parameters. By selecting other functions instead (2.2) and (2.4) one would end with less restrictive SCC. The model which was employed here indicates the importance of the valence $q \bar q$ structure \cite{17} in the meson state. There is some hope that such relativistic model, at least in some richer version, can play a useful role in the classification of meson states. It can be a useful tool in the design of future experiments if it can provide a reasonable estimate of the magnitude of expected experimental effects. \newpage \begin{Large}
\section{introduction} Hawking's discovery of black hole evaporation posed the question about the final fate of a black hole. This issue is far from being fully understood but exploiting two-dimensional (2d) models gained insight into this intriguing puzzle. In particular, Bose, Parker and Peleg (BPP) suggested the exactly solvable modification of the Callan-Giddings-Harvey-Strominger (CGHS) model \cite{callan} in which a black hole evaporates leaving, as an end state, a regular static semi-infinite throat \cite{bose}. The existence of such a type of solution motivates an interest to finding regular static solutions in 2d dilaton gravity in a more general setting. Apart from the possible role of such solutions as ''remnants'' after the evaporation of black hole, this task is also motivated by the necessity to elucidate the structure of exactly solvable solutions (of any types) in dilaton gravity. In the paper \cite{kaz} the approach was suggested which was based on a treatment of non linear $\sigma $ model. A more simple approach appealing directly to the properties of the action coefficient as functions of a dilaton field was proposed in \cite{zasl99} where it was also shown that the general structure of found exactly solvable models encompasses all known particular ones. The paper \cite{zasl99} deals only with black holes with a regular horizon. Correspondingly, the temperature is put equal to the Hawking one. In the present paper we consider another types of a metric and find the general form of static solutions in exactly solvable models at an arbitrary temperature. We carry out the analysis of the obtained general solutions for the BPP model and generalizations of it which, as well as BPP\ ones, do not contain singularities at finite values of a dilaton field. It turns out that in these models there exists a rather rich family of solutions regular everywhere. In spite of the term ''throat''\ in 1+1 dimensional world is somewhat conditional, we will use it, following \cite{bose}, to denote any regular geodesically complete geometry extending to infinity (which for definiteness is supposed to be a left one) where its curvature is nonzero. If a metric is flat there, we will speak about a soliton-like configuration. It is supposed that at the right infinity a spacetime is flat in any case. We will see that, depending on temperature, the found family of exact solutions contains regular spacetimes of all three possible types - black holes, throats or soliton-like configurations. \section{general form of static solutions} Consider the action \begin{equation} I=I_{0}+I_{PL}\text{,} \label{action} \end{equation} where \begin{equation} I_{0}=\frac{1}{2\pi }\int_{M}d^{2}x\sqrt{-g}[F(\phi )R+V(\phi )(\nabla \phi )^{2}+U(\phi )] \label{clac} \end{equation} and the Polyakov-Liouville action \cite{pl} incorporating effects of Hawking radiation and its back reaction on the black hole metric can be written as \begin{equation} I_{PL}=-\frac{\kappa }{2\pi }\int_{M}d^{2}x\sqrt{-g}[\frac{(\nabla \psi )^{2}% }{2}+\psi R] \label{PL} \end{equation} The function $\psi $ obeys the equation \begin{equation} \Box \psi =R \label{psai} \end{equation} where $\Box =\nabla _{\mu }\nabla ^{\mu }$, $\kappa =N/24$ is the quantum coupling parameter, $N$ is number of scalar massless fields, $R$ is a Riemann curvature. We omit the boundary terms in the action as we are interested only in field equations and their solutions. Varying the action with respect to a metric gives us $(T_{\mu \nu }=2\frac{% \delta I}{\delta g^{\mu \nu }})$: \begin{equation} T_{\mu \nu }\equiv T_{\mu \nu }^{(0)}+T_{\mu \nu }^{(PL)}=0 \label{6} \end{equation} where \begin{equation} T_{\mu \nu }^{(0)}=\frac{1}{2\pi }\{2(g_{\mu \nu }\Box F-\nabla _{\mu }\nabla _{\nu }F)-Ug_{\mu \nu }+2V\nabla _{\mu }\phi \nabla _{\nu }\phi -g_{\mu \nu }V(\nabla \phi )^{2}\}\text{,} \label{7} \end{equation} \begin{equation} T_{\mu \nu }^{(PL)}=-\frac{\kappa }{2\pi }\{\partial _{\mu }\psi \partial _{\nu }\psi -2\nabla _{\mu }\nabla _{\nu }\psi +g_{\mu \nu }[2R-\frac{1}{2}% (\nabla \psi )^{2}]\} \label{8} \end{equation} Variation of the action with respect to $\phi $ gives rise to the equation \begin{equation} R\frac{dF}{d\phi }+\frac{dU}{d\phi }=2V\Box \phi +\frac{dV}{d\phi }(\nabla \phi )^{2} \label{9} \end{equation} Hereafter we assume that a dilaton field is not constant identically (such special kinds of solutions are considered in \cite{solod}, \cite{zasl98}). Then, as is shown in \cite{kaz}, \cite{zasl99}, we can gain a rich set of exactly solvable models if the action coefficients in (\ref{clac}) $F(\phi )$% , $V(\phi )$ and $U(\phi )\equiv 4\lambda ^{2}e^{\int_{0}^{\phi }\omega (\phi ^{\prime })d\phi ^{\prime }}$are restricted by one constraint equation. This equation can be written as \begin{equation} V=\omega (u-\frac{\kappa \omega }{2}) \label{v} \end{equation} where $u\equiv \frac{dF}{d\phi }$. (In fact, eq. (\ref{v}) admits further generalization due to a possible term $A(u-\kappa \omega )^{2}$ but, as it leads, generally speaking, to metrics which are not asymptotically flat, hereafter we put $A=0$). Then it turns out that the metric is a pure static. In the present paper we will use the conformal gauge \begin{equation} ds^{2}=g(-dt^{2}+d\sigma ^{2}) \label{conf} \end{equation} where $g=g(\sigma )$. In this gauge the curvature $R=-g^{-1}(g^{\prime }/g)^{\prime }$. Throughout the paper the prime denotes differentiation with respect to a spatial coordinate. Substituting it into (\ref{psai}) we get after simple manipulations \begin{equation} g=e^{-\psi -a\sigma } \label{g} \end{equation} where $a$ is a constant. It is implied that time is normalized in such a way that in (\ref{g}) the function $g\rightarrow 1$ at $\sigma \rightarrow \infty $. After simple rearrangement the (00) and (11) field equations (\ref{6}) with the metric (\ref{g}) are reduced to one equation \begin{equation} \xi _{1}\phi ^{\prime \prime }+\xi _{2}\phi ^{\prime 2}-\xi _{1}\frac{% g^{\prime }\phi ^{\prime }}{g}=0 \label{field} \end{equation} where $\xi _{1}=\frac{d\tilde{F}}{d\phi }$, $\xi _{2}=\frac{d^{2}\tilde{F}}{% d\phi ^{2}}-\tilde{V}$, $\tilde{F}=F-\kappa \psi $, $\tilde{V}=V-\frac{% \kappa }{2}(\frac{d\psi }{d\phi })^{2}$. In \cite{zasl99} we found that for exactly solvable models (\ref{v}) the function $\psi $ is equal to $\psi \equiv \psi _{0}(\phi )$ where \begin{equation} \psi _{0}=\int \omega d\phi \label{ps0} \end{equation} For a given metric this function is defined up to an arbitrary function $% \chi $ such that $\Box \chi =0$ as it follows from (\ref{psai}). For a static metric (\ref{conf}) this gives us (up to the constant) $\chi =\gamma \sigma $ where $\gamma $ is a constant. We will see below that for $\gamma =0 $ we return to black hole solutions obtained in \cite{zasl99} whereas for $\gamma \neq 0$ we find new types of static solutions which include, in particular, finite temperature generalizations of ''semi-infinite throats'' \cite{bose}. In what follows we will deal with solutions of the form \begin{equation} \psi =\psi _{0}+\gamma \sigma \label{ps} \end{equation} with an arbitrary $\gamma \neq 0$. At the right infinity where spacetime is supposed to be flat the function $\psi \sim \sigma \rightarrow \infty $ for a generic $g$ \cite{sol95} irrespectively of whether or not $\gamma =0$. Meanwhile, at the left coordinate infinity $\sigma \rightarrow -\infty $ (a reader should bear in mind that if a spacetime is not geodesically complete, the coordinate $\sigma $ does not cover the whole manifold) the admittance of $\gamma \neq 0$ can qualitatively change the character of solution. For instance, instead of a black hole for which at a horizon $g\rightarrow 0$ and $\psi $ is finite \cite{solod} one can get at a left infinity a ''semi-infinite throat'' \cite{bose} for which both $g$ and $\psi $ diverge there to ensure, by definition of the ''semi-infinite throat'', the divergencies of a proper distance at a finite value of a coordinate $\sigma $ (see below for details). The structure of the action coefficients in our case is the same as in \cite {zasl99}. Namely, the relationship between them is represented by eq. (\ref {v}). The new feature introduced in our treatment as compared to \cite {zasl99} is connected with another character of boundary conditions for the function $\psi .$ Whereas in \cite{solod}, \cite{zasl99} one deals with a black hole in the Hartle-Hawking state for which $\psi =\psi _{0}$ is regular on a horizon, for solutions discussed below $\psi $ may diverge at $% \sigma \rightarrow -\infty $ due to the term $\gamma \sigma $ in (\ref{ps}). It is convenient to split coefficients in eq. (\ref{field}) into two parts singling out the term which is built up with the help of $\psi _{0}$: $\xi _{1}=\xi _{1}^{(0)}-\kappa \gamma \frac{d\sigma }{d\phi }$, $\xi _{2}=\xi _{2}^{(0)}-\kappa \gamma \frac{d^{2}\sigma }{d\phi ^{2}}+\kappa [\frac{d\psi _{0}}{d\phi }\gamma \frac{d\sigma }{d\phi }+\frac{1}{2}(\gamma \frac{d\sigma }{d\phi })^{2}]$, \begin{equation} \xi _{1}^{(0)}=\frac{d\tilde{F}^{(0)}}{d\phi },\xi _{2}^{(0)}=\frac{d^{2}% \tilde{F}^{(0)}}{d\phi ^{2}}-\tilde{V}^{(0)},\tilde{F}^{(0)}=F-\kappa \psi _{0},\tilde{V}^{(0)}=V-\frac{\kappa }{2}(\frac{d\psi _{0}}{d\phi })^{2}. \label{0} \end{equation} Then eq. (\ref{field}) takes the form \begin{equation} \xi _{1}^{(0)}\phi ^{\prime \prime }+\xi _{2}^{(0)}\phi ^{\prime 2}+\xi _{1}^{(0)}\phi ^{\prime }(\psi ^{(0)^{\prime }}+\delta )=\kappa \gamma (\delta -\frac{\gamma }{2}) \label{f2} \end{equation} where $\delta =\gamma +a$. Let us multiply this equation by the factor $\eta $ such that $\xi _{2}^{(0)}\eta =\frac{d(\xi _{1}^{(0)}\eta )}{d\phi }$. Then eq. (\ref{f2}) can be cast into the form \begin{equation} z^{\prime }+z(\psi ^{(0)^{\prime }}+\delta )=\kappa \gamma (\delta -\frac{% \gamma }{2})\eta \label{y} \end{equation} where $z=\eta \xi _{1}^{(0)}\phi ^{\prime }=\eta \frac{d\tilde{F}}{d\sigma }% ^{(0)}$. It follows from (\ref{v}), (\ref{ps0}) and (\ref{0}) that $\eta =e^{-\psi _{0}}$. Then after integration we get from (\ref{y}) \begin{equation} \tilde{F}^{(0)}=C+De^{-\sigma \delta }+\kappa \gamma (1-\frac{\gamma }{% 2\delta })\sigma \label{F} \end{equation} It follows from the trace of eq. (\ref{7}), (\ref{8}) that \begin{equation} U=\Box \tilde{F}=\Box \tilde{F}^{(0)} \label{pot} \end{equation} As for any function $f(\sigma )$ we have in the metric (\ref{g}) $\Box f=g^{-1}f^{\prime \prime }$, we get from (\ref{pot}), (\ref{g}) and (\ref {ps0}) the relationship between constants $D\delta ^{2}=4\lambda ^{2}$. In what follows we assume that the function $\tilde{F}$ as well as $F$ has the asymptotic behavior $\tilde{F}\simeq e^{\omega _{0}\phi }$ ($\omega _{0}=const$) at the right infinity $\sigma \rightarrow \infty $ where $\phi \rightarrow -\infty $, $\psi _{0}\simeq \omega _{0}\phi $ and the metric is flat, $g\simeq e^{-\omega _{0}\phi -\delta \sigma }\rightarrow 1$. This asymptotic condition is satisfied, in particular, by CGHS \cite{callan}, RST \cite{rst} and BPP models with $\omega _{0}=-2$. Thus, we get $D=1$, $\delta =-2\lambda $ (it is assumed for definiteness that $\omega _{0}<0$) and the relationship between a dilaton field and coordinate reads $\tilde{F}% ^{(0)}=f(y)\equiv e^{2y}-By+C$ where $y\equiv \lambda \sigma $, $B=-\kappa \frac{\gamma }{\lambda }(1+\frac{\gamma }{4\lambda })$. The value of $\gamma $ can be found from asymptotical conditions at right infinity where spacetime is supposed to be flat. Let us impose such a condition which describes quantum fields at finite temperature $T$. It means that at right infinity the quantum stress-energy tensor should have the form \begin{mathletters} \begin{equation} T_{\mu }^{\nu (PL)}=\frac{\pi ^{2}N}{6}T^{2}(1,-1) \label{T} \end{equation} Comparing it with the asymptotic form of eq. (\ref{8}), we find that $T_{\mu }^{\nu (PL)}=\frac{\kappa }{4\pi }\psi _{\infty }^{\prime 2}(1,-1)$ where $% \psi _{\infty }^{\prime }=\lim_{\sigma \rightarrow \infty }\frac{d\psi }{% d\sigma }$. Hence, $\psi _{\infty }^{\prime 2}=\frac{2\pi ^{2}T^{2}N}{% 3\kappa }=16\pi ^{2}T^{2}$. Remembering that for exactly solvable models under consideration the Hawking temperature is equal to $T_{0}=\lambda /2\pi $ \cite{zasl99} we obtain $\psi _{\infty }^{\prime 2}=4\lambda ^{2}\frac{% T^{2}}{T_{0}^{2}}$. On the other hand, we have from (\ref{g}) and the condition $g\rightarrow 1$ at infinity that $\psi _{\infty }^{\prime }=-a$. Then $\gamma =\delta -a=-2\lambda +\psi _{\infty }^{\prime }$ $=2\lambda (T/T_{0}-1)$ whence $B=\kappa (1-T^{2}/T_{0}^{2})$ (the sign of $\psi _{\infty }^{\prime }$ is chosen to ensure $\gamma =0$ for $T=T_{0}$ when $% \psi =\psi _{0}$ and we return to the situation described in \cite{zasl99}). Collecting all basic formulas, we obtain \end{mathletters} \begin{eqnarray} ds^{2} &=&g(-dt^{2}+d\sigma ^{2})\text{, }g=\exp (-\psi _{0}+2y)\text{, }% y=\lambda \sigma \text{, }\psi _{0}=\int \omega d\phi \text{.} \label{bas} \\ \tilde{F}^{(0)}(\phi ) &=&f(y)\equiv e^{2y}-By+C\text{, }B=\kappa (1-T^{2}/T_{0}^{2})\text{, }T_{0}=\lambda /2\pi \text{.} \nonumber \end{eqnarray} If $T>T_{0}$, the coefficient $B>0$ and the function takes its minimum at the point $y_{0}=\frac{1}{2}\ln B$ where $f(y_{0})=C-C^{*}$, \begin{equation} C^{*}=-\frac{B}{2}(1+\ln \frac{2}{B}) \label{c} \end{equation} Let $T=T_{0}$, then the coefficients $\gamma =B=0$ whence we obtain from (% \ref{bas}) $g=\exp (-\psi _{0})(\tilde{F}^{(0)}-C)$. If we identify $C=% \tilde{F}(\phi _{h})$ we obtain black hole solutions with the horizon at $% \phi =\phi _{h}$ analyzed in \cite{zasl99}. Then the equality of temperatures simply means that the state of the Euclidean metric we deal with is the Hartle-Hawking one. To make the comparison more clear, let us pass from the conformal gauge (\ref{g}) to the Schwarzschild one $% ds^{2}=-gdt^{2}+g^{-1}dx^{2}$used in \cite{zasl99}. This can be achieved by the coordinate transformation $x=\int d\sigma g$. It follows from (\ref{bas}% ) that $x=(2\lambda )^{-1}\int d\phi e^{-\psi }\tilde{F}^{\prime }$ that agrees with eq. (27) of \cite{zasl99}. It is worth stressing, however, that the class of solutions (\ref{bas}) is much more general than that found in \cite{zasl99}. It includes not only black holes due to the parameter $B$ which describes a system at arbitrary temperature when black holes with a regular horizon cannot exist. In what follows we will use the notation $T_{0}=\lambda /2\pi $ even in cases when $% T_{0}$ does not have the meaning of the Hawking temperature. Equations (\ref {bas}) represent in the closed form the general expression describing static solutions in exactly solvable models of 2d gravity \cite{kaz}, \cite{zasl99}% . The function in the right hand side of eq. (\ref{bas}) has the universal form while a particular model is characterized by the choice of $\tilde{F}% ^{(0)}(\phi )$. It is the analysis of this class of solutions that we now turn to. As we are mainly interested in finding regular solutions, we discuss the BPP model and its generalizations. \section{BPP model at finite temperature} This model is specified by the choice \begin{equation} F=e^{-2\phi }-2\kappa \phi \text{, }\tilde{F}=e^{-2\phi }\text{, }\omega =-2% \text{, }\psi _{0}=-2\phi \text{, }V=4e^{-2\phi }+2\kappa \label{bpp} \end{equation} If $T=T_{0}$ we have $B=0$ and $e^{2y}=e^{-2\phi }-e^{-2\phi _{h}}$, $% g=1-e^{2(\phi -\phi _{h})}$. As seen from the above formulas, the expression for the metric (either in terms of $\phi $ or in terms of a coordinate) has the same form as in the classical limit ($\kappa =0$), so in this sense the metric does not acquire quantum corrections \cite{zasl99}. In the paper \cite {bose} the authors considered the case $T=0$. Then asymptotically the energy density of quantum field approaches zero (radiationless solutions). In general, spacetime contains a time-like singularity or a singular horizon depending on the value of $C$ that determines whether or not the function $% f(y)$ has a zero. The special case arises when simultaneously $% f(y)=0=f^{\prime }(y)$ at $y=y_{0}$, $C=C^{*}=-\frac{\kappa }{2}(1+\ln \frac{% 2}{\kappa })$ \cite{bose} (a reader should have in mind that we define the quantum parameter $\kappa =N/24$ whereas in \cite{bose} $\kappa =N/12$). Then the spacetime does not contain singularities at all extending from the left infinity where $R=-4\lambda ^{2}$ (semi-infinite throat) to the right one where $R=0$. In fact, the analysis is tractable to the case of an arbitrary $T$ due to the universality and simplicity of the function $f(y)$ with an arbitrary $B$% . Let, first, $T<T_{0}$, so $B>0$. For $C<C^{*}$ there is a time-like naked singularity $(y=y_{0}$) at a finite proper distance. If $C>C^{*}$, there exists a singular horizon at $\phi =\infty $. The throat exists provided $% C=C^{*}$ where $C^{*}$, according to (\ref{c}), depends on temperature via $% B(T)$. The calculation of curvature shows that at the throat $R=-4\lambda ^{2}$ independently of temperature. In the limit $T=0$ $B=\kappa $ and we return to radiationless solutions described in \cite{bose}. If $T>T_{0}$, $B<0$ the function $f(y)$ is monotonic and does not turn into a zero. Then for any $C$ there is a time-like singularity at $\phi =\infty $. In the point $T=T_{0}$ the coefficients $B=0=C^{*}$. If $C>0$ we have a black hole in the Hartle-Hawking state with a regular horizon. If $C<0$ there exists a time-like singularity. For $C=0$ the spacetime is flat (linear dilaton vacuum $\phi =-y$). The situation can be summarized as follows. BPP model \begin{tabular}{|l|l|l|l|} \hline & $T<T_{0}$ & $T=T_{0}$, $B=0=C^{*}$ & $T>T_{0}$ \\ \hline $C<C^{*}$ & time-like singularity & time-like singularity & time-like \\ \hline $C=C^{*}$ & throat & linear dilaton vacuum, flat spacetime & singularity \\ \hline $C>C^{*}$ & singular horizon & space-like singularity behind a horizon & for any $C$ \\ \hline \end{tabular} The main feature of the model in question is the fact that the semi-infinite throat may exist at any finite temperature $T<T_{0}$. Meanwhile, this needs the special choice $C=C^{*}(T)$. It turns out, however, that there are another exactly solvable models for which the existence of a throat is generic and does need fine tuning in constants. \section{generalizations of bpp model} Consider the model \begin{eqnarray} F^{(0)} &=&\alpha e^{-2b\phi }+e^{-2\phi }-2\kappa \phi \text{, }\alpha >0% \text{, }0<b\leq 1/2\text{, }\omega =-2\text{, }V=4(e^{-2\phi }+b\alpha e^{-\phi })+2\kappa \text{,} \label{gen} \\ \tilde{F}^{(0)} &=&\alpha e^{-2b\phi }+e^{-2\phi } \nonumber \end{eqnarray} which is, according to (\ref{v}), exactly solvable and the Hawking temperature for black hole solutions is equal to $T_{0}$ \cite{zasl99}. Let first $B>0$ ($T<T_{0}$). Then near the point $y_{0}$ where $f(y_{0})=0$ we have for any $C<C^{*}$ $f\sim y-y_{0}$, $\tilde{F}^{(0)}\sim e^{-2b\phi }\rightarrow 0$, $\phi \rightarrow \infty $, so $e^{2\phi }$ $\sim (y-y_{0})^{-1/b}$ and, according to (\ref{bas}), $g=e^{2\phi +2y}\sim (y-y_{0})^{-1/b}$. Then the curvature $R\sim (y-y_{0})^{1/b-2}$. If $b<1/2$, the curvature $R=0$ at left infinity, the proper distance $l$ between $y_{0}$ and any other point $y>y_{0}$ diverges like $l\sim (y-y_{0})^{1-1/2b}$. Thus, we obtain a soliton-like configuration. If $C=C^{*}$, $f\sim (y-y_{0})^{2}$, $g$ $\sim (y-y_{0})^{-2/b}$, $R\sim (y-y_{0})^{2/b-2}\rightarrow 0$, so the configuration is again soliton-like for any $0<b<1$. For $C>C^{*}$ spacetime cannot be regular everywhere, it contains the singular horizon where $y\rightarrow -\infty $, $g\sim e^{2y}\left| y\right| ^{-1/b}$, $R\sim -e^{-2y}\left| y\right| ^{1/b-2}$, a proper distance $l$ is finite. The rest of cases can be treated in a similar manner. We only dwell on the fact that for $T=T_{0}$ a spacetime represent a black hole whose curvature is finite everywhere. In particular, at left infinity where $\phi \rightarrow \infty $ the curvature $R\rightarrow 0$, so spacetime is asymptotically flat at both infinities. The variety of cases is tabulated as follows. $\tilde{F}^{(0)}=e^{-2\phi }+\alpha e^{-2b\phi }$, $0<b<1/2$ \begin{tabular}{|l|l|l|l|} \hline & $T<T_{0}$ & $T=T_{0}$, $B=0=C^{*}$ & $T>T_{0}$ \\ \hline $C<C^{*}$ & soliton-like & soliton-like & soliton-like \\ \hline $C=C^{*}$ & soliton-like for $0<b<1$ & soliton-like & for \\ \hline $C>C^{*}$ & singular horizon & black hole regular everywhere & any $C$ \\ \hline \end{tabular} Let now $b=1/2$. The results of consideration, details of which we omit, can be summarized in the table below. $\tilde{F}^{(0)}=e^{-2\phi }+\alpha e^{-\phi }$ \begin{tabular}{|l|l|l|l|} \hline & $T<T_{0}$ & $T=T_{0}$, $B=0=C^{*}$ & $T>T_{0}$ \\ \hline $C<C^{*}$ & throat & throat & throat \\ \hline $C=C^{*}$ & soliton-like & soliton-like & for \\ \hline $C>C^{*}$ & singular horizon & black hole regular everywhere & any $C$ \\ \hline \end{tabular} The most interesting feature of this model is the existence of a throat at any temperature (for $C<C^{*}$, if $T\leq T_{0}$ and for an arbitrary $C$ if $T>T_{0}$). \section{Discussion and conclusion} We have found a general form of static solutions in 2d exactly solvable dilaton gravity theories accounting for quantum effects at finite temperature. The criteria of solvability are the same as derived in \cite {kaz}, \cite{zasl99} for eternal black holes and represent one constraint on the action coefficient (\ref{v}). Thus, every model with black hole solutions considered in \cite{zasl99} generates a family of solutions of another kind which generalize \cite{zasl99} and reduce to it in the particular case $T=T_{0}$ when the solution describes a black hole with a Hawking temperature $T_{0}$ in thermal equilibrium with its Hawking radiation. It was indicated in \cite{zasl99} that exactly solvable models in question exhibits a series of universal properties. In particular, it is remarkable that a Hawking temperature is constant for all such models, $% T_{0}=\lambda /2\pi $. The present consideration extended this universality to a much more wide class solution and enabled us to classified them in an unified manner due to the universal structure of (\ref{bas}) independently of the choice of the model. In fact, the particular properties of the model are encoded in the only function $\tilde{F}^{(0)}(\phi )$. If $\tilde{F}% ^{(0)\prime }(\phi )$ does not have zeros at finite $\phi $, as it takes place for the BPP model \cite{bose} and its generalization considered in the present paper, the only relevant information is contained in the asymptotic behavior of $\tilde{F}^{(0)}$ at $\phi \rightarrow \infty $. We analyzed two concrete examples and found that there exist three types of solutions regular everywhere. First, it is a black hole, if $T=T_{0}$ (the possibility of solutions of such a kind was pointed out in \cite{zasl99}). Second, this a semi-infinite throat generalizing the observation made in \cite{bose}. The most interesting feature of throats found in the model (\ref{gen}) with $% b=1/2$ is that such types of solutions, unlike the BPP case, do not need a special choice of parameters of the solution:\ semi-infinite throats may exist in the whole range of parameters and temperatures. Third case is a soliton-like configuration extending in both directions to flat infinities without horizons. Such a type of solution is absent in the BPP model but may occur in the generalizations of it. The conditions which make the existence of semi-infinite throats possible are essentially quantum. Indeed, as seen from (\ref{bas}), such a kind of configuration arises only if the function $f(y)$ is not monotonic due to the coefficient $B$ and takes its minimum value at some finite $y_{0}=\frac{1}{2}% \ln B$. However, in the classical limit $\kappa \rightarrow 0$ the coefficient $B\rightarrow 0$ and we get $g=e^{-\psi _{0}}(F^{(0)}-C)$. Choosing $C=F(\phi _{h})$ we can get a classical counterpart of a quantum black hole but not a semi-infinite throat. For example, with $\omega =-2$, $% F^{(0)}=e^{-2\phi }$ we have the familiar expression for a static dilaton black hole $g=1-Ce^{2\phi }$ where $C/2$ plays the role of a classical ADM mass. In the present paper we determined possible types of static solutions but did not consider dynamical scenarios in the process of which they could arise. Here we only restrict ourselves by a remark that finite temperature solutions allow us to consider more complicated processes of interaction between a black hole and its quantum surrounding than a evaporation into vacuum. In particular, this enables us to include into consideration directly the role of thermal bath \cite{bath}. Another interested problem for further researches is looking for possibilities to gain new exactly solvable models not only due to relaxing the regularity condition for the function $\psi $ at the left infinity (as was made in the present paper) but also to the generalization of the action structure itself. All this needs a special treatment. I\ am grateful to Sergey Solodukhin for careful reading the manuscript and valuable comments. This work is supported by International Science Education Program (ISEP), grant No. QSU082068.
\section{An invariant condition for CR--maps} Assume that $M\subset\mathbb{C}^{N}$ and $M^\prime\subset \mathbb{C}^{N^\prime}$ are generic real--analytic submanifolds of $\mathbb{C}^{N}$ and $\mathbb{C}^{N^\prime}$, respectively. We will consider CR--maps $H:M\to M^\prime$, that is, maps which push complex tangent vectors to $M$ to complex tangent vectors to $M^\prime$, ie, $H_{*}T^c_p M\subset T^c_{H(p)} M^\prime$. For our purposes, this means that after choosing coordinates $(Z_1^\prime,\dots, Z_{N^\prime}^\prime)\in\mathbb{C}^{N^\prime}$ the components to $H$ in this coordinate system are CR--functions on $M$ (For details concerning CR--maps and the last statement, the reader is encouraged to see \cite{BERbook}, \S 2.3.). We want to give a condition on the map $H$ which guarantees that it is the restriction of a holomorphic map from $\mathbb{C}^{N}$ into $\mathbb{C}^{N^\prime}$. Our nondegeneracy condition involves taking derivatives of the complex gradients to $M^\prime$ pulled back to $M$ via $H$ with respect to the CR--vector fields tangent to $M$. For simplicity, assume that $0\in M$, $0\in M^\prime$ and $H(0)= 0$ (in the examples, this will not always be the case). Let $\rho^\prime = (\rho^\prime_1,\dots,\rho^\prime_{d^\prime})$ be a local defining function for $M^\prime$ near $0\in\mathbb{C}^{N^\prime}$. We write $\crb{M}$ for the CR--bundle of $M$. \begin{defin} \label{D:nondeg} Assume that $H$ is a CR--mapping of $M$ into $M^\prime$ of class $C^{k_0}$ and $H(0)=0$. Define complex linear subspaces $E_k (0)\subset\mathbb{C}^{N^\prime}$ by \begin{multline} \label{E:ek} E_k(0) = \spanc \{ L_1 \cdots L_j \, \rho^\prime_{s Z^\prime} \left(H(Z),\overline{H(Z)}\right)\Bigr|_{Z=0} \colon L_r \in \Gamma( M ,\crb{M}) \\ \text{ for } 1\leq r \leq j , 0\leq j \leq k, 1\leq l\leq d^\prime \}. \end{multline} We say that $H$ is $k_0$--nondegenerate if $E_{k_0}(0) = \mathbb{C}^{N^\prime}$ and $E_k(0) \neq \mathbb{C}^{N^\prime}$ for $k < k_0$. Here we have used the notation $\rho^\prime_{s Z^\prime} = ( \rho^\prime_{s Z^\prime_1},\dots, \rho^\prime_{s Z^\prime_{N^\prime}} )$ for the complex gradient. \end{defin} We will show that Definition~\ref{D:nondeg} is invariant under biholomorphic changes of coordinates in Lemma \ref{L:ind}. \begin{rem} If there exists a $k_0$--nondegenerate map into the manifold $M^\prime$, then automatically $M^\prime$ is finitely nondegenerate. Here we say that a real submanifold $M$ is finitely nondegenerate (or more specifically $k_0$--nondegenerate) if the identity mapping of $M$ is $k_0$--nondegenerate in the sense of Definition~\ref{D:nondeg}. For the original definition of finite nondegeneracy, see \cite{BERbook}, \S 11.1.. In fact, if there is a $k_0$--nondegenerate map $H$ into $M^\prime$, then $M^\prime$ will be $\ell_0$--nondegenerate with $k_0\geq\ell_0$. The reader can check this by writing out the condition in Definition \ref{D:nondeg} in coordinates and using the chain rule. \end{rem} Let $\Gamma$ be an open, convex cone in $\mathbb{R}^{d}$. By a wedge $\hat W_\Gamma\subset\mathbb{C}^{N}$with edge $M$ we mean a set of the form $\hat W_\Gamma=\{Z\in U \colon \rho(Z,\bar Z ) \in \Gamma \}$ where $\rho$ is a defining function for $M$ and $U\subset\mathbb{C}^{N}$ is an open neighbourhood of the origin. We can now state our result. \begin{thm} \label{T:main} Let $M\subset\mathbb{C}^{N}$, $M^\prime\subset\mathbb{C}^{N^\prime}$ be generic real--analytic submanifolds, $0\in M$, $0\in M^\prime$. Assume that $H:M\to M^\prime$ is a CR--map with $H(0)=0$ which is $k_0$--nondegenerate at $0$ and extends continuously to a holomorphic function (called again $H$) in a wedge $\hat W_\Gamma$ with edge $M$. Then $H$ extends to a holomorphic map $H:\mathbb{C}^{N}\to\mathbb{C}^{N^\prime}$ in a full neighbourhood of $0\in\mathbb{C}^{N}$. \end{thm} Let us note that if we assume $M$ to be of finite type, then by a result of Tumanov \cite{TU1} every CR--map $H$ will automatically extend to a wedge. Also, the hypotheses that $H$ extends continuously and that $H$ is $C^{k_0}$ on $M$ implies that actually $H$ extends as a $C^{k_0}$ function up to the edge $M$ of $\hat W_\Gamma$ (see e.g. \cite{BERbook}, \S 7.5.). Note that if $M\subset\mathbb{C}^{N}$, $M^\prime\subset\mathbb{C}^{N}$ are of equal codimension and both $\ell_0$--nondegenerate, then every CR--diffeomorphism of class $\ell_0$ is $\ell_0$--nondegenerate. Let us now illustrate Definition \ref{D:nondeg} by giving some examples. \begin{ex}\label{EX:first} Let $M=\{(z,w)\in\mathbb{C}^2 \colon \imag w = |z|^4 \} $, $M^\prime = \{(z,w)\in\mathbb{C}^2 \colon \imag w = |z|^2 \}$. Then the map $H(z,w)=(z^2,w)$ is $2$--nondegenerate. \end{ex} We will now give an example where $M$ has codimension $1$ and $M^\prime$ has codimension $2$. \begin{ex} Consider the manifolds $M=\{(z,w)\in\mathbb{C}^2 \colon \imag w = |z|^2 \}$, $M^\prime = \{ (z^\prime,w_1^\prime,w_2^\prime)\colon \imag w_1^\prime = \imag w_2^\prime = |z^\prime |^2 \}$. The map $H:\mathbb{C}^2\to\mathbb{C}^3$, $H(z,w) = (z,w,w)$ is $1$--nondegenerate.\end{ex} \begin{ex} As the source manifold consider the ball in $\mathbb{C}^2$, $M= \{ (z_1, z_2)\in \mathbb{C}^2 \colon |z_1|^2 + |z_2|^2 = 1 \}$. There are (up to automorphisms) exactly three nonlinear maps from $M$ into the $3$--ball $M^\prime = \{ (z_1^\prime, z_2^\prime, z_3^\prime)\in \mathbb{C}^3 \colon |z_1^\prime|^2+ |z_2^\prime|^2 + |z_3^\prime|^2 = 1 \}$, as Faran \cite{FA1} showed: $H_1(z_1,z_2) = (z_1,z_1 z_2, z_2^2)$, $H_2(z_1,z_2) = (z_1^2,\sqrt{2}z_1 z_2, z_2^2)$, and $H_3(z_1,z_2) = (z_1^3,\sqrt{3}z_1 z_2, z_2^3)$. These are $2$, $2$, and $3$--nondegenerate, respectively, at the point $(1,0,0)$. However, if we consider maps from an $n$--ball into an $n+1$--ball which are at least $C^3$ for $n\geq 3$, a result due to Webster \cite{WE1} tells us that these are all linear and hence never nondegenerate. \end{ex} We would like to note that variants of Definition \ref{D:nondeg} have appeared in the literature before, for example in the case of pseudoconvex hypersurfaces (\cite{CSK}, \cite{FA2}) but the invariant formulation given here is new. Another similar condition can be found in \cite{HAN}. In this paper Han studies the reflection principle for Levi--nondegenerate hypersurfaces. He requires that after putting the target hypersurface in suitable coordinates $(z^\prime,w^\prime)$ and writing $H=(f,g)$ in these coordinates, that the derivatives $L^\alpha \bar f (0)$ where $L^\alpha = L_1^{\alpha_1}\cdots L_n^{\alpha_n}$ and $L_1,\dots,L_n$ is a basis for the CR--vector fields tangent to $M$ span $\mathbb{C}^{n^\prime}$ (where $N^\prime= n^\prime+1$). The problem with this condition is that it is not invariant under biholomorphic changes of coordinates, and Theorem 1.1 in \cite{HAN} is incorrect, as the following example which is due to D. Zaitsev shows. \begin{ex}\label{EX:dime} Consider the hypersurface $M^\prime \subset \mathbb{C}^3$ given by $\imag w = |z_1|^2 - |z_2|^2$. After the change of coordinates $z_1 = \zeta_1 +\zeta_2 - \zeta_2^2$, $z_2 = \zeta_2$, $w = \tau$, $M$ takes the form $\imag \tau = |\zeta_1 +\zeta_2 -\zeta_2^2|^2 - |\zeta_2 |^2$. Let $\phi$ be a CR--function on $M= \{ (z,w) \colon \imag w = |z|^2 \}$ and consider the map $H(z,w)= (\phi^2 (z,w), \phi (z,w), 0)$.Then $f = (\phi^2,\phi)$ and if we choose $\phi$ such that $\phi_z(0)\neq 0$, then the derivatives $L \bar f(0)$ and $L^2 \bar f (0)$ will span $\mathbb{C}^2$. Choosing for $\phi$ any CR--function which does not extend (such exist since $M$ is strongly pseudoconvex) gives a counterexample to Theorem 1.1 in \cite{HAN}. \end{ex} We would now like to mention some of the reflection principles which have been obtained for manifolds in complex spaces of different dimensions. In the case of pseudoconvex hypersurfaces, reflection principles were obtained in \cite{FO1} and \cite{HU1}. A reflection principle for balls has been obtained in \cite{BHR}. Also, the reader is referred to the survey article \cite{FO2} for some other special cases which have been settled. We will conclude this section by proving that Definition~\ref{D:nondeg} is independent of the choice of coordinates. The definition is independent of the choice of defining function, as the reader can easily prove by induction on $k$. The definition is also independent of the choice of coordinates $Z\in\mathbb{C}^{N}$. It only remains to be shown that it is independent of changes of holomorphic coordinates in the target space. \begin{lem}\label{L:ind} Let $M\subset\mathbb{C}^{N}$, $M^\prime\subset\mathbb{C}^{N^\prime}$, $H$ be as in Definition \ref{D:nondeg}. Define $E_k(0)$ by \eqref{E:ek}. If we change coordinates by $\tilde Z^\prime = F(Z^\prime)$ in $\mathbb{C}^{N^\prime}$, and denote the corresponding spaces defined by \eqref{E:ek} in the new coordinates by $\tilde E_k(0)$, then \begin{equation}\label{E:ektransform} \tilde E_k (0) = E_k (0) \left( \vardop{F}{Z^\prime}(0)\right)^{-1}, \end{equation} where $E_k(0)$ is considered as a space of row vectors. \end{lem} \begin{proof} Suppose that $F:\mathbb{C}^{N^\prime}\to\mathbb{C}^{N^\prime}$ is a local biholomorphic change of coordinates with $F(0)=0$. Since Definition~\ref{D:nondeg} is independent of the choice of defining function we can choose $\tilde\rho = \rho^\prime \circ F^{-1}$ as a defining function for $F(M)$. By the chain rule we have that \[\tilde \rho_{ j \tilde Z^\prime}= (\rho^\prime_j \circ F^{-1})_{\tilde Z^\prime} = (\rho^\prime_{j Z^\prime} \circ F^{-1}) \left( \vardop{F^{-1}}{\tilde{Z}^\prime} \right). \] Now the only crucial point is that the matrix on the right hand side has holomorphic entries. Hence, if we compose with $H$, this matrix will be annihilated by the CR--vector fields on $M$. Applying the vector fields $ L_1,\dots, L_r$ as in \eqref{E:ek} to this equation we conclude that \begin{multline} L_1\cdots L_r \tilde \rho_{ j \tilde Z^\prime} (F\circ H(Z),\overline{F\circ H(Z)}) \\= L_1 \cdots L_r \rho^\prime_{j Z^\prime} (H(Z),\overline{H(Z)}) \left( \vardop{F^{-1}}{\tilde{Z}}(F(H(Z))) \right). \end{multline} Evaluating this identity at $0$, we obtain \eqref{E:ektransform}. \end{proof} \section{Preliminaries}\label{S:prel} Before we give the proof of Theorem~\ref{T:main}, we want to give the technical details on which the proof is based. We will be making use of the Edge--of--the--Wedge Theorem. The technique used here is basically the same as in \cite{BJT}. We start by fixing a defining equation for $M$. Assume that $\codim M = d$. Then it is possible to choose coordinates $(z,w)\in \mathbb{C}^{n}\times\mathbb{C}^{d} = \mathbb{C}^{N}$ near $0$ and a real--analytic function $\varphi:\mathbb{C}^{n}\times\mathbb{R}^{d} \to \mathbb{R}^{d}$, defined near $0$, such that $M$ is given by \begin{equation}\label{E:definingequM} \imag w = \varphi (z,\bar z, \real w ), \end{equation} and $\varphi(0)=0$, $d\varphi (0)=0$, $\varphi (z, 0,s) = \varphi (0,\bar z,s) = 0$. Define the function $\Psi : \mathbb{C}^{n}\times\mathbb{R}^{d}\times\mathbb{R}^{d} \to \mathbb{C}^{N}$, \[ \Psi(z,\bar z,s,t) = (z,s+it + i\varphi (z,\bar z, s+ it) ). \] This is a real--analytic diffeomorphism, but {\em not} holomorphic. It flattens out $M$ in the sense that near the origin $\Psi^{-1} (M) = \mathbb{C}^{n}\times\mathbb{R}^{d}\times \{0\}$. With respect to $s+it$, that is, for $z$ fixed, $\Psi$ is actually holomorphic. We shall equip $\mathbb{C}^{n}\times\mathbb{R}^{d}$ with the CR--structure transported by $\Psi$ from $M$. Let $\Gamma$ be an open, convex cone in $\mathbb{R}^{d}$ (all cones will be assumed to be open and convex). By a wedge $W_\Gamma$ we mean a set of the form \[ W_\Gamma = \{ (z,s,t) \in \mathbb{C}^{n}\times \mathbb{R}^{d}\times\mathbb{R}^{d} \colon t\in\Gamma \}. \] We will also call the image under $\Psi$ of $W_\Gamma$ a wedge, but to avoid confusion we will write $\tilde W_\Gamma$ for it. We will also write $W_\Gamma^\epsilon = \{ (z,s,t)\in W_\Gamma \colon |z|<\epsilon, |s|<\epsilon, |t| < \epsilon \}$ and $\tilde W_\Gamma^\epsilon$ for the image under $\Psi$ of $W_\Gamma^\epsilon$. The apparent difference with the definition of the wedge $\hat W_\Gamma$ given before the statement of Theorem~\ref{T:main} is reconciled by the observation that given any cone $\Gamma^\prime$ whose closure is contained in $\Gamma$, then in the sense of germs of sets, $\hat W_{\Gamma^\prime} \subset \tilde W_\Gamma$ and $ \tilde W_{\Gamma^\prime} \subset \hat W_\Gamma $ (see the end of Section 2 of \cite{BJT}). By saying that a CR--function $h$ on $M$ extends holomorphically to $W_\Gamma^\epsilon$ we shall mean that there is a holomorphic function $f$ on $\tilde W_\Gamma^\epsilon$, continuous up to $M$, such that $h(z,\bar z, s) = f(\Psi(z,\bar z,s,0))$, and likewise we will say that $h$ extends holomorphically to a full neighbourhood of the origin if there is a holomorphic function $f$ defined in a full neighbourhood of the origin such that this equality holds. To say that $h$ extends holomorphically to a full neighbourhood of the origin is the same as having an $\epsilon>0$ such that for each $z$ with $|z|<\epsilon$ the function $s\mapsto h(z,\bar z,s)$ extends holomorphically in $s+it$ for $|s|<\epsilon$, $|t|<\epsilon$ such that the function $h(z,\bar z, s,t)$ is a bounded, measurable function in all of its variables for $|z|<\epsilon, |s|<\epsilon, |t|<\epsilon$ (for details, see e.g. \cite{BERbook}, \S 1.7.). The Edge--of--the--Wedge theorem ensures that this will be the case if $h$ extends holomorphically in $s+it$ for each fixed $z$ to $W_\Gamma^\epsilon$ and $W_{-\Gamma}^\epsilon$ (for details, see Section 2 of \cite{BJT}) in a way such that $h$ is a continuous function of all its variables in $W_\Gamma^\epsilon\cup W_{-\Gamma}^\epsilon\cup (M\cap \{|z|<\epsilon, |s|<\epsilon\})$. Actually, this assumption might seem unnecessarily strong to the reader; in fact, it is enough to assume that the extension has boundary value in the sense of distributions $h$, but a regularity theorem implies that if $h$ is continuous on $M$, then this extension is actually continuous in the sense above (see e.g. \cite{BERbook}, \S 7.3. and \cite{Rosay}). Since we shall only deal with functions which are at least continuous on $M$, we shall assume continuity of the extension a priori, and shall not go into further detail. For the following arguments, we will fix $\Gamma$ and say that $h$ {\it extends up} differentiably (respectively continuously) if it extends holomorphically to $W_\Gamma^\epsilon$ in a way such that the extended function is a $C^1$ (respectively continuous) function of all of its variables in $\bar W_\Gamma^\epsilon=W_\Gamma^\epsilon \cup M$ (where $M$ is shrunk appropriately) and that $h$ {\it extends down} if it extends holomorphically to $\bar W_{-\Gamma}^\epsilon$ in a way such that the extended function is a $C^1$ (respectively continuous) function of all of its variables (this is the terminology used in \cite{BERbook}, \S 9.2., in the case of hypersurfaces). The main point can be phrased as follows: \begin{center} {\em If $h$ extends up continuously then $\bar h$ extends down continuously.} \end{center} In fact, the extension of $\bar{h}$ is just given by $\bar h (z,\bar z, s, -t) = \overline{h(z,\bar z, s, t)}$. This will be used later on for the components of the CR--mapping $H$. We also need to know how $Lh$ behaves for vector fields $L$ tangent to $M$ whose coefficients extend up continuously. Here the result is that if the coefficients of $L$ extend up continuously and $h$ extends up differentiably, then $Lh$ extends up continuously. Similarily, if the coefficients of $L$ extend up smoothly, the regularity of the extension is dropped by $1$ (that is, if $h$ extends up in a $C^k$ manner, $Lh$ does so in a $C^{k-1}$ manner). We summarize the discussion: \begin{lem}\label{L:prel} Let $h$ be a CR--function on $M$. If $h$ extends up continuously (respectively differentiably), then $\bar h$ extends down continuously (respectively differentiably). If $L$ is a vector field tangent to $M$ whose coefficients extend up smoothly and $h$ extends up of order $C^k$ then $Lh$ extends up of order $C^{k-1}$. If $h$ extends up and down continuously, then $h$ extends holomorphically to a full neighbourhood of the origin. \end{lem} The next lemma is used in order to actually calculate with Definition~\ref{D:nondeg}. Its proof is an easy induction on $k$ which is left to the reader. \begin{lem} \label{L:ekinbasis} Let $L_1,\dots , L_n$ be a local basis near $0$ of $\Gamma(M,\crb{M})$. For every multiindex $\alpha=(\alpha_1,\dots,\alpha_n)\in\mathbb{N}^{n}$ define ${L}^\alpha = {L_1}^{\alpha_1} \cdots{L_n}^{\alpha_n}$. Then \begin{equation} \label{E:ekinbasis} E_k(0) = \spanc \{ {L}^\alpha \rho^\prime_{l Z^\prime}(H(Z),\overline{H(Z)} ) \big|_{Z=0} \colon 1\leq l\leq d^\prime , |\alpha |\leq k \}, \end{equation} where the $E_k(0)$ are defined by \eqref{E:ek}. \end{lem} \section{Proof of the Reflection Principle} Choose coordinates $(Z_1^\prime,\dots,Z_{N^\prime}^\prime)\in\mathbb{C}^{N^\prime}$ and a real--analytic defining function $\rho^\prime$ for $M^\prime$ defined in some neighbourhood of $0$. In these coordinates, $H=(H_1,\dots,H_{N^\prime})$ with each $H_j$ a CR--function on $M$. We will think of $M$ as $\mathbb{C}^{n}\times\mathbb{R}^{d}$ as explained in Section~\ref{S:prel}. A local basis of the CR--vector fields (for which we will use the vector notation and write $\Lambda=(\Lambda_1,\dots,\Lambda_n$)) is then given in matrix notation by (with $\varphi$ as in \eqref{E:definingequM}; we also refer the reader to \cite{BERbook}, \S 1.6.) \begin{equation} \label{E:thelambdas} \Lambda = \vardop{}{\bar z} - i (\varphi_{\bar z})^T (I + i \varphi_s)^{-1} \vardop{}{s}. \end{equation} The coefficients of all of these vector fields clearly extend up and down (smoothly), since $\varphi$ is real--analytic. Since $H$ maps $M$ into $M^\prime$, $\rho^\prime (H(Z),\overline{H(Z)}) = 0$ for $Z\in M$. Applying $\Lambda_1,\dots,\Lambda_n$ repeatedly and using the chain rule we see that for every multiindex $\alpha\in\mathbb{N}^{n}$, $|\alpha|\leq k_0$ (as in Lemma~\ref{L:ekinbasis}, $\Lambda^\alpha= \Lambda^{\alpha_1}\cdots\Lambda^{\alpha_n}$) and for every $l$, $1\leq l \leq d^\prime$, \begin{equation}\label{E:phialpha} 0 = \Lambda^\alpha \rho^\prime_l (H(Z),\overline{H(Z)}) = \Phi_{l \alpha} (H(Z),\overline{H(Z)}, (\Lambda^\beta \overline{H(Z)})_{1\leq |\beta |\leq k_0} ), \quad Z\in M, \end{equation} where $\Phi_{l\alpha}$ is a real--analytic function defined and convergent on a neighbourhood of $\{0\}\times\{0\}\times \mathbb{C}^{K(k_0)}\subset\mathbb{C}^{N^\prime}\times \mathbb{C}^{N^\prime}\times\mathbb{C}^{K(k_0)}$, $K(k_0)$ being the cardinality of the set $\{ \beta\in\mathbb{N}^{n}\colon 1\leq |\beta | \leq k_0\}$. By our assumption and Lemma~\ref{L:ekinbasis} we can choose $N^\prime$ multiindices $\alpha^1,\dots,\alpha^{N^\prime}$, $0\leq |\alpha^j |\leq k_0$ and integers $l^1,\dots,l^{N^\prime}$, $1\leq l^j \leq d^\prime$, such that \begin{equation}\label{E:fullspan} \spanc \{ \Lambda^{\alpha^j} \rho^\prime_{l^j Z^\prime} (H(Z),\overline{H(Z)}) \big|_{Z=0} \colon 1\leq j \leq N^\prime \} = \mathbb{C}^{N^\prime}. \end{equation} We consider the system of equations \begin{equation}\label{E:system} \Phi_{l^j \alpha^j} (X_1,\dots,X_{N^\prime},Y_1, \dots,Y_{N^\prime},W)=0,\quad 1\leq j\leq N^\prime, \end{equation} where $W\in\mathbb{C}^{K(k_0)}$. We claim that \eqref{E:system} admits a (unique) real--analytic solution in $(X_1,\dots,X_{N^\prime})$ in a neighbourhood of the point $(0,0,(\Lambda^\beta \bar H (0))_{1\leq |\beta |\leq k_0})$. In fact, if we compute the Jacobian of this system with respect to $X_1,\dots, X_{N^\prime}$ it is of full rank at this point because of \eqref{E:fullspan}. So we can invoke the implicit function theorem to conclude that there are real--analytic functions $\Upsilon_1,\dots,\Upsilon_{N^\prime}$, convergent on a neighbourhood $U$ of $(0,(\Lambda^\beta \bar H (0))_{1\leq |\beta |\leq k_0})$, such that the unique solution of \eqref{E:system} in $U$ is given by \begin{equation}\label{E:sol} X_j = \Upsilon_j (Y_1,\dots,Y_{N^\prime},W), \quad 1\leq j \leq N^\prime. \end{equation} Recalling \eqref{E:phialpha} we conclude that \begin{equation}\label{E:solinH} H_j (Z) = \Upsilon_j (\overline{H_1 (Z)}, \dots, \overline{H_{N^\prime} (Z)},(\Lambda^\beta \overline{H(Z)} )_{1\leq |\beta |\leq k_0 } ). \end{equation} Now the proof is finished by using Lemma~\ref{L:prel}: Each $H_j$ is assumed to extend up; so each $\bar{H}_j$ extends down (of order $C^{k_0}$, by a regularity theorem, see Theorem 7.5.1. in \cite{BERbook}). Hence the whole right hand side of \eqref{E:solinH} extends down continuously (after choosing $\epsilon$ small enough). This shows that each $H_j$ extends down continuously. Since each $H_j$ also extends up, Lemma~\ref{L:prel} implies that each $H_j$ extends holomorphically to a full neighbourhood of the origin. The theorem is proved. \bibliographystyle{plain}
\section{Introduction: microcanonical description of finite systems} In a microcanonical description of finite systems, the magnetic equation of state and the zero-field isothermal magnetic susceptibility are defined in terms of derivatives of the microcanonical entropy $s({\scriptstyle \varepsilon,m,L^{-1}})$ (cf.~\cite{KPH98},\cite{FN1}):\\ \begin{equation} \label{mmic} m_{\scriptscriptstyle {h=0}} ( {\scriptstyle \varepsilon,L^{-1}} ) \quad \Longleftarrow \quad \max_{m\ge0}\left\{ s({\scriptstyle \varepsilon,m,L^{-1}} ) \right\}\,\,=\,\, s\left( {\scriptstyle \varepsilon,m_{\scriptscriptstyle {h=0}} (\varepsilon,L^{-1}),L^{-1}} \right) \end{equation} \begin{eqnarray} \label{chimic} \lefteqn{ \chi_{\scriptscriptstyle T;{h=0}}({\scriptstyle \varepsilon , L^{-1} } ) \,\, := \,\, \left. \frac{\partial m}{\partial h} \right|_{T;{h=0}}( {\scriptstyle \varepsilon,L^{-1}} )\,\,=} \\ & & =\,\, \left\{\left[ \frac{\partial s}{\partial \varepsilon} \left[ \left( \frac{\partial^2 s}{\partial \varepsilon \partial m}\right)^2\left/ \frac{\partial^2 s}{\partial \varepsilon^2}\right. -\frac{\partial^2s}{\partial m^2} \right]^{-1}\right]( {\scriptstyle \varepsilon,m,L^{-1}} )\right\}_{m=m_{\scriptscriptstyle {h=0}} (\varepsilon,L^{-1})} \nonumber \end{eqnarray} Here, $\varepsilon$ denotes the specific enthalpy (cf.~\cite{FN0}), $h$ the magnetic field, $m$ the specific magnetization, $L$ the linear system size and $d$ the spatial dimension. For notational convenience, the zero field limit is denoted by $h=0$. By definition, the microcanonical entropy is the logarithm of the microcanonical partition function (density of states) $\Omega$: \begin{equation} s({\scriptstyle \varepsilon,m,L^{-1}} ) \,=\, L^{-d}\,\ln \Omega ({\scriptstyle \varepsilon,m,L^{-1}} )\;\;\;. \label{smic} \end{equation} In Fig.~1, the zero-field magnetization (Fig.~1a) and the zero-field magnetic susceptibility (Fig.~1b) are plotted for the example of the $3d$-Ising system for various system sizes. Both observables show a behaviour reminiscent of the critical behaviour of the infinite system and therefore suggest the introduction of finite system exponents $\beta_{\varepsilon,L}$ and $\gamma_{\varepsilon,L}$ (describing the power law behaviour with respect to the base $|\tilde{\varepsilon} |$ defined below; cf.~Ref.~13 of \cite{KPH98}).\\ The transition energies $\varepsilon_T(L)$, at which the non-analyticities occur, are obviously dependent on system size.\\ In Figs.~1c/d, magnetizations and susceptibilities are plotted as functions of the reduced enthalpy $\tilde{\varepsilon}(L) := (\varepsilon-\varepsilon_T(L))/|\varepsilon_T(L)|$. \begin{center} \begin{figure}[h] \setlength{\unitlength}{0.1bp} \special{! /gnudict 120 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /userlinewidth gnulinewidth def /vshift -33 def /dl {10 mul} def /hpt_ 31.5 def /vpt_ 31.5 def /hpt hpt_ def /vpt vpt_ def /M {moveto} bind def /L {lineto} bind def /R {rmoveto} bind def /V {rlineto} bind def /vpt2 vpt 2 mul def /hpt2 hpt 2 mul def /Lshow { currentpoint stroke M 0 vshift R show } def /Rshow { currentpoint stroke M dup stringwidth pop neg vshift R show } def /Cshow { currentpoint stroke M dup stringwidth pop -2 div vshift R show } def /UP { dup vpt_ mul /vpt exch def hpt_ mul /hpt exch def /hpt2 hpt 2 mul def /vpt2 vpt 2 mul def } def /DL { Color {setrgbcolor Solid {pop []} if 0 setdash } {pop pop pop Solid {pop []} if 0 setdash} ifelse } def /BL { stroke gnulinewidth 2 mul setlinewidth } def /AL { stroke gnulinewidth 2 div setlinewidth } def /UL { gnulinewidth mul /userlinewidth exch def } def /PL { stroke userlinewidth setlinewidth } def /LTb { BL [] 0 0 0 DL } def /LTa { AL [1 dl 2 dl] 0 setdash 0 0 0 setrgbcolor } def /LT0 { PL [] 0 1 0 DL } def /LT1 { PL [4 dl 2 dl] 0 0 1 DL } def /LT2 { PL [2 dl 3 dl] 1 0 0 DL } def /LT3 { PL [1 dl 1.5 dl] 1 0 1 DL } def /LT4 { PL [5 dl 2 dl 1 dl 2 dl] 0 1 1 DL } def /LT5 { PL [4 dl 3 dl 1 dl 3 dl] 1 1 0 DL } def /LT6 { PL [2 dl 2 dl 2 dl 4 dl] 0 0 0 DL } def /LT7 { PL [2 dl 2 dl 2 dl 2 dl 2 dl 4 dl] 1 0.3 0 DL } def /LT8 { PL [2 dl 2 dl 2 dl 2 dl 2 dl 2 dl 2 dl 4 dl] 0.5 0.5 0.5 DL } def /Pnt { stroke [] 0 setdash gsave 1 setlinecap M 0 0 V stroke grestore } def /Dia { stroke [] 0 setdash 2 copy vpt add M hpt neg vpt neg V hpt vpt neg V hpt vpt V hpt neg vpt V closepath stroke Pnt } def /Pls { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /Box { stroke [] 0 setdash 2 copy exch hpt sub exch vpt add M 0 vpt2 neg V hpt2 0 V 0 vpt2 V hpt2 neg 0 V closepath stroke Pnt } def /Crs { stroke [] 0 setdash exch hpt sub exch vpt add M hpt2 vpt2 neg V currentpoint stroke M hpt2 neg 0 R hpt2 vpt2 V stroke } def /TriU { stroke [] 0 setdash 2 copy vpt 1.12 mul add M hpt neg vpt -1.62 mul V hpt 2 mul 0 V hpt neg vpt 1.62 mul V closepath stroke Pnt } def /Star { 2 copy Pls Crs } def /BoxF { stroke [] 0 setdash exch hpt sub exch vpt add M 0 vpt2 neg V hpt2 0 V 0 vpt2 V hpt2 neg 0 V closepath fill } def /TriUF { stroke [] 0 setdash vpt 1.12 mul add M hpt neg vpt -1.62 mul V hpt 2 mul 0 V hpt neg vpt 1.62 mul V closepath fill } def /TriD { stroke [] 0 setdash 2 copy vpt 1.12 mul sub M hpt neg vpt 1.62 mul V hpt 2 mul 0 V hpt neg vpt -1.62 mul V closepath stroke Pnt } def /TriDF { stroke [] 0 setdash vpt 1.12 mul sub M hpt neg vpt 1.62 mul V hpt 2 mul 0 V hpt neg vpt -1.62 mul V closepath fill} def /DiaF { stroke [] 0 setdash vpt add M hpt neg vpt neg V hpt vpt neg V hpt vpt V hpt neg vpt V closepath fill } def /Pent { stroke [] 0 setdash 2 copy gsave translate 0 hpt M 4 {72 rotate 0 hpt L} repeat closepath stroke grestore Pnt } def /PentF { stroke [] 0 setdash gsave translate 0 hpt M 4 {72 rotate 0 hpt L} repeat closepath fill grestore } def /Circle { stroke [] 0 setdash 2 copy hpt 0 360 arc stroke Pnt } def /CircleF { stroke [] 0 setdash hpt 0 360 arc fill } def /C0 { BL [] 0 setdash 2 copy moveto vpt 90 450 arc } bind def /C1 { BL [] 0 setdash 2 copy moveto 2 copy vpt 0 90 arc closepath fill vpt 0 360 arc closepath } bind def /C2 { BL [] 0 setdash 2 copy moveto 2 copy vpt 90 180 arc closepath fill vpt 0 360 arc closepath } bind def /C3 { BL [] 0 setdash 2 copy moveto 2 copy vpt 0 180 arc closepath fill vpt 0 360 arc closepath } bind def /C4 { BL [] 0 setdash 2 copy moveto 2 copy vpt 180 270 arc closepath fill vpt 0 360 arc closepath } bind def /C5 { BL [] 0 setdash 2 copy moveto 2 copy vpt 0 90 arc 2 copy moveto 2 copy vpt 180 270 arc closepath fill vpt 0 360 arc } bind def /C6 { BL [] 0 setdash 2 copy moveto 2 copy vpt 90 270 arc closepath fill vpt 0 360 arc closepath } bind def /C7 { BL [] 0 setdash 2 copy moveto 2 copy vpt 0 270 arc closepath fill vpt 0 360 arc closepath } bind def /C8 { BL [] 0 setdash 2 copy moveto 2 copy vpt 270 360 arc closepath fill vpt 0 360 arc closepath } bind def /C9 { BL [] 0 setdash 2 copy moveto 2 copy vpt 270 450 arc closepath fill vpt 0 360 arc closepath } bind def /C10 { BL [] 0 setdash 2 copy 2 copy moveto vpt 270 360 arc closepath fill 2 copy moveto 2 copy vpt 90 180 arc closepath fill vpt 0 360 arc closepath } bind def /C11 { BL [] 0 setdash 2 copy moveto 2 copy vpt 0 180 arc closepath fill 2 copy moveto 2 copy vpt 270 360 arc closepath fill vpt 0 360 arc closepath } bind def /C12 { BL [] 0 setdash 2 copy moveto 2 copy vpt 180 360 arc closepath fill vpt 0 360 arc closepath } bind def /C13 { BL [] 0 setdash 2 copy moveto 2 copy vpt 0 90 arc closepath fill 2 copy moveto 2 copy vpt 180 360 arc closepath fill vpt 0 360 arc closepath } bind def /C14 { BL [] 0 setdash 2 copy moveto 2 copy vpt 90 360 arc closepath fill vpt 0 360 arc } bind def /C15 { BL [] 0 setdash 2 copy vpt 0 360 arc closepath fill vpt 0 360 arc closepath } bind def /Rec { newpath 4 2 roll moveto 1 index 0 rlineto 0 exch rlineto neg 0 rlineto closepath } bind def /Square { dup Rec } bind def /Bsquare { vpt sub exch vpt sub exch vpt2 Square } bind def /S0 { BL [] 0 setdash 2 copy moveto 0 vpt rlineto BL Bsquare } bind def /S1 { BL [] 0 setdash 2 copy vpt Square fill Bsquare } bind def /S2 { BL [] 0 setdash 2 copy exch vpt sub exch vpt Square fill Bsquare } bind def /S3 { BL [] 0 setdash 2 copy exch vpt sub exch vpt2 vpt Rec fill Bsquare } bind def /S4 { BL [] 0 setdash 2 copy exch vpt sub exch vpt sub vpt Square fill Bsquare } bind def /S5 { BL [] 0 setdash 2 copy 2 copy vpt Square fill exch vpt sub exch vpt sub vpt Square fill Bsquare } bind def /S6 { BL [] 0 setdash 2 copy exch vpt sub exch vpt sub vpt vpt2 Rec fill Bsquare } bind def /S7 { BL [] 0 setdash 2 copy exch vpt sub exch vpt sub vpt vpt2 Rec fill 2 copy vpt Square fill Bsquare } bind def /S8 { BL [] 0 setdash 2 copy vpt sub vpt Square fill Bsquare } bind def /S9 { BL [] 0 setdash 2 copy vpt sub vpt vpt2 Rec fill Bsquare } bind def /S10 { BL [] 0 setdash 2 copy vpt sub vpt Square fill 2 copy exch vpt sub exch vpt Square fill Bsquare } bind def /S11 { BL [] 0 setdash 2 copy vpt sub vpt Square fill 2 copy exch vpt sub exch vpt2 vpt Rec fill Bsquare } bind def /S12 { BL [] 0 setdash 2 copy exch vpt sub exch vpt sub vpt2 vpt Rec fill Bsquare } bind def /S13 { BL [] 0 setdash 2 copy exch vpt sub exch vpt sub vpt2 vpt Rec fill 2 copy vpt Square fill Bsquare } bind def /S14 { BL [] 0 setdash 2 copy exch vpt sub exch vpt sub vpt2 vpt Rec fill 2 copy exch vpt sub exch vpt Square fill Bsquare } bind def /S15 { BL [] 0 setdash 2 copy Bsquare fill Bsquare } bind def /D0 { gsave translate 45 rotate 0 0 S0 stroke grestore } bind def /D1 { gsave translate 45 rotate 0 0 S1 stroke grestore } bind def /D2 { gsave translate 45 rotate 0 0 S2 stroke grestore } bind def /D3 { gsave translate 45 rotate 0 0 S3 stroke grestore } bind def /D4 { gsave translate 45 rotate 0 0 S4 stroke grestore } bind def /D5 { gsave translate 45 rotate 0 0 S5 stroke grestore } bind def /D6 { gsave translate 45 rotate 0 0 S6 stroke grestore } bind def /D7 { gsave translate 45 rotate 0 0 S7 stroke grestore } bind def /D8 { gsave translate 45 rotate 0 0 S8 stroke grestore } bind def /D9 { gsave translate 45 rotate 0 0 S9 stroke grestore } bind def /D10 { gsave translate 45 rotate 0 0 S10 stroke grestore } bind def /D11 { gsave translate 45 rotate 0 0 S11 stroke grestore } bind def /D12 { gsave translate 45 rotate 0 0 S12 stroke grestore } bind def /D13 { gsave translate 45 rotate 0 0 S13 stroke grestore } bind def /D14 { gsave translate 45 rotate 0 0 S14 stroke grestore } bind def /D15 { gsave translate 45 rotate 0 0 S15 stroke grestore } bind def /DiaE { stroke [] 0 setdash vpt add M hpt neg vpt neg V hpt vpt neg V hpt vpt V hpt neg vpt V closepath stroke } def /BoxE { stroke [] 0 setdash exch hpt sub exch vpt add M 0 vpt2 neg V hpt2 0 V 0 vpt2 V hpt2 neg 0 V closepath stroke } def /TriUE { stroke [] 0 setdash vpt 1.12 mul add M hpt neg vpt -1.62 mul V hpt 2 mul 0 V hpt neg vpt 1.62 mul V closepath stroke } def /TriDE { stroke [] 0 setdash vpt 1.12 mul sub M hpt neg vpt 1.62 mul V hpt 2 mul 0 V hpt neg vpt -1.62 mul V closepath stroke } def /PentE { stroke [] 0 setdash gsave translate 0 hpt M 4 {72 rotate 0 hpt L} repeat closepath stroke grestore } def /CircE { stroke [] 0 setdash hpt 0 360 arc stroke } def /BoxFill { gsave Rec 1 setgray fill grestore } def end } \begin{picture}(1800,1611)(300,0) \special{" gnudict begin gsave 0 0 translate 0.100 0.100 scale 0 setgray newpath LTb 350 250 M 18 0 V 1382 0 R -18 0 V 350 515 M 18 0 V 1382 0 R -18 0 V 350 780 M 18 0 V 1382 0 R -18 0 V 350 1044 M 18 0 V 1382 0 R -18 0 V 350 250 M 0 18 V 0 993 R 0 -18 V 817 250 M 0 18 V 0 993 R 0 -18 V 1283 250 M 0 18 V 0 993 R 0 -18 V 1750 250 M 0 18 V 0 993 R 0 -18 V LTb 350 250 M 1400 0 V 0 1011 V -1400 0 V 350 250 L 1.000 UL LT0 723 780 M 163 0 V 1750 250 M -19 0 V -18 0 V -19 0 V -19 0 V -18 69 V -19 109 V -19 36 V -18 39 V -19 36 V -19 21 V -18 32 V -19 27 V -19 22 V -18 22 V -19 19 V -19 19 V -18 21 V -19 19 V -19 16 V -18 17 V -19 15 V -19 17 V -18 15 V -19 15 V -19 14 V -18 15 V -19 13 V -19 13 V -18 14 V -19 12 V -19 12 V -18 13 V -19 11 V -19 14 V -18 11 V -19 10 V -19 12 V -18 11 V -19 13 V -19 10 V -18 10 V -19 12 V -19 7 V -18 12 V -19 10 V -19 9 V -18 10 V -19 9 V -19 10 V -18 9 V -19 11 V -19 10 V -18 10 V -19 7 V -19 8 V -18 9 V -19 9 V -19 9 V -18 8 V -19 10 V -19 9 V -18 7 V -19 7 V 1.000 UL LT1 723 680 M 163 0 V 1750 250 M -4 0 V -11 0 V -11 0 V -11 0 V -11 0 V -10 0 V -11 0 V -11 0 V -11 0 V -11 0 V -10 0 V -11 0 V -11 0 V -11 0 V -11 0 V -10 0 V -11 0 V -11 0 V -11 0 V -11 0 V -10 0 V -11 0 V -11 0 V -11 37 V -11 77 V -10 53 V -11 25 V -11 27 V -11 28 V -11 22 V -10 16 V -11 22 V -11 15 V -11 16 V -11 18 V -10 12 V -11 15 V -11 16 V -11 11 V -11 13 V -10 14 V -11 9 V -11 13 V -11 12 V -11 10 V -10 10 V -11 10 V -11 12 V -11 11 V -11 9 V -10 10 V -11 10 V -11 8 V -11 10 V -11 9 V -10 9 V -11 10 V -11 8 V -11 8 V -11 9 V -10 8 V -11 8 V -11 8 V -11 9 V -11 7 V -10 8 V -11 8 V -11 7 V -11 7 V -11 9 V -10 7 V -11 7 V -11 6 V -11 9 V -11 6 V -11 6 V -10 7 V -11 8 V -11 6 V -11 7 V -11 5 V -10 9 V -11 4 V -11 6 V -11 8 V -11 7 V -10 4 V -11 6 V -11 7 V -11 7 V -11 6 V -10 6 V -11 4 V -11 5 V -11 8 V -11 7 V -10 3 V 1.000 UL LT2 723 580 M 163 0 V 1750 250 M -1 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 0 V -7 0 V -6 0 V -7 0 V -7 0 V -7 56 V -7 39 V -7 24 V -6 29 V -7 25 V -7 19 V -7 20 V -7 16 V -6 13 V -7 17 V -7 12 V -7 15 V -7 12 V -6 12 V -7 13 V -7 8 V -7 11 V -7 11 V -6 7 V -7 11 V -7 8 V -7 11 V -7 9 V -6 9 V -7 9 V -7 9 V -7 7 V -7 8 V -6 8 V -7 8 V -7 7 V -7 7 V -7 7 V -6 9 V -7 7 V -7 6 V -7 7 V -7 6 V -6 8 V -7 7 V -7 6 V -7 6 V -7 7 V -6 5 V -7 7 V -7 6 V -7 6 V -7 6 V -6 6 V -7 6 V -7 6 V -7 6 V -7 6 V -6 5 V -7 5 V -7 6 V -7 5 V -7 6 V -6 5 V -7 5 V -7 6 V -7 6 V -7 5 V -6 4 V -7 6 V -7 5 V -7 4 V -7 5 V -6 5 V -7 5 V -7 5 V -7 5 V -7 4 V -7 5 V -6 5 V -7 4 V -7 5 V -7 5 V -7 4 V -6 4 V -7 5 V -7 5 V -7 4 V -7 5 V -6 4 V -7 4 V -7 4 V -7 5 V -7 4 V -6 4 V -7 5 V -7 4 V -7 4 V -7 4 V -6 4 V -7 4 V -7 4 V -7 5 V -7 3 V -6 4 V -7 5 V -7 3 V -7 4 V -7 4 V -6 4 V -7 4 V -7 4 V -7 4 V -7 3 V -6 4 V -7 4 V -7 4 V -7 3 V -7 4 V -6 4 V -7 3 V -7 4 V -7 3 V -7 4 V -6 4 V -7 4 V -7 3 V -7 4 V -7 3 V -6 3 V -7 4 V -7 4 V -7 3 V -7 3 V -6 4 V -7 3 V -7 4 V -7 3 V -7 3 V -6 4 V -7 3 V -7 3 V -7 4 V -7 3 V -6 3 V -7 4 V -7 3 V -7 3 V -7 3 V -6 4 V -7 3 V -7 3 V -7 3 V -7 3 V -7 4 V -6 3 V -7 3 V -7 3 V 1.000 UL LT3 723 480 M 163 0 V 1750 250 M -4 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -5 0 V -5 0 V -4 0 V -5 49 V -4 29 V -5 19 V -4 36 V -5 13 V -4 21 V -5 18 V -5 7 V -4 15 V -5 12 V -4 12 V -5 12 V -4 10 V -5 7 V -5 11 V -4 9 V -5 9 V -4 9 V -5 7 V -4 8 V -5 6 V -4 10 V -5 7 V -5 8 V -4 7 V -5 4 V -4 9 V -5 7 V -4 7 V -5 5 V -4 8 V -5 4 V -5 7 V -4 7 V -5 5 V -4 6 V -5 6 V -4 5 V -5 6 V -4 6 V -5 5 V -5 6 V -4 4 V -5 6 V -4 5 V -5 3 V -4 7 V -5 6 V -4 4 V -5 6 V -5 4 V -4 5 V -5 5 V -4 5 V -5 4 V -4 5 V -5 4 V -4 6 V -5 4 V -5 4 V -4 4 V -5 5 V -4 4 V -5 4 V -4 4 V -5 5 V -4 4 V -5 4 V -5 5 V -4 4 V -5 3 V -4 4 V -5 5 V -4 3 V -5 5 V -4 3 V -5 3 V -5 6 V -4 4 V -5 2 V -4 5 V -5 4 V -4 3 V -5 3 V -5 5 V -4 3 V -5 4 V -4 3 V -5 4 V -4 3 V -5 4 V -4 4 V -5 3 V -5 4 V -4 3 V -5 4 V -4 3 V -5 3 V -4 2 V -5 6 V -4 2 V -5 4 V -5 4 V -4 2 V -5 4 V -4 3 V -5 3 V -4 4 V -5 3 V -4 4 V -5 1 V -5 4 V 1.000 UL LT4 723 380 M 163 0 V 1604 250 M -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 0 V -4 74 V -3 7 V -3 5 V -3 29 V -3 20 V -4 13 V -3 11 V -3 8 V -3 10 V -3 13 V -4 8 V -3 10 V -3 6 V -3 12 V -3 7 V -4 9 V -3 6 V -3 7 V -3 7 V -3 6 V -4 7 V -3 8 V -3 4 V -3 8 V -3 7 V -4 4 V -3 5 V -3 5 V -3 7 V -3 7 V -4 6 V -3 2 V -3 7 V -3 5 V -3 5 V -4 5 V -3 5 V -3 5 V -3 5 V -3 4 V -4 5 V -3 5 V -3 4 V -3 5 V -3 3 V -4 5 V -3 5 V -3 3 V -3 5 V -3 5 V -4 5 V -3 4 V -3 2 V -3 5 V -3 4 V -4 3 V -3 4 V -3 6 V -3 2 V -3 3 V -4 5 V -3 4 V -3 3 V -3 4 V -3 4 V -4 4 V -3 3 V -3 4 V -3 5 V -3 1 V -4 4 V -3 4 V -3 3 V -3 3 V -3 4 V -4 3 V -3 3 V -3 4 V -3 3 V -3 3 V -4 4 V -3 3 V -3 4 V -3 2 V -3 4 V -4 3 V -3 3 V -3 2 V -3 5 V -4 2 V -3 3 V -3 3 V -3 4 V -3 2 V -4 3 V -3 3 V -3 3 V -3 3 V -3 3 V -4 3 V -3 3 V -3 4 V -3 2 V -3 2 V -4 3 V -3 3 V -3 3 V -3 2 V -3 3 V -4 3 V -3 3 V -3 3 V -3 3 V -3 2 V -4 3 V -3 3 V -3 2 V -3 2 V -3 3 V -4 3 V -3 2 V -3 3 V -3 2 V -3 2 V -4 4 V -3 2 V -3 3 V -3 1 V -3 4 V -4 3 V -3 1 V -3 4 V -3 2 V -3 2 V -4 2 V -3 3 V -3 2 V -3 3 V -3 3 V -4 1 V -3 4 V -3 -2 V -3 5 V -3 3 V -4 4 V -3 1 V -3 3 V -3 1 V -3 3 V -4 3 V -3 2 V -3 1 V -3 4 V -3 2 V -4 2 V -3 4 V -3 1 V -3 1 V -3 3 V -4 2 V -3 3 V -3 1 V -3 3 V -3 2 V -4 3 V -3 1 V -3 3 V -3 1 V stroke grestore end showpage } \put(673,380){\makebox(0,0)[r]{$\scriptstyle L=18$}} \put(673,480){\makebox(0,0)[r]{$\scriptstyle L=16$}} \put(673,580){\makebox(0,0)[r]{$\scriptstyle L=14$}} \put(673,680){\makebox(0,0)[r]{$\scriptstyle L=12$}} \put(673,780){\makebox(0,0)[r]{$\scriptstyle L=10$}} \put(1050,1411){\makebox(0,0){{\bf a)} $\qquad m_{\scriptscriptstyle {h=0}}({\scriptstyle \varepsilon,L^{-1}})\,\,$ {\em vs.} $\varepsilon$\rule{19mm}{0mm}}} \put(1750,150){\makebox(0,0){-0.8}} \put(1283,150){\makebox(0,0){-0.9}} \put(817,150){\makebox(0,0){-1}} \put(350,150){\makebox(0,0){-1.1}} \put(300,1044){\makebox(0,0)[r]{0.3}} \put(300,780){\makebox(0,0)[r]{0.2}} \put(300,515){\makebox(0,0)[r]{0.1}} \put(300,250){\makebox(0,0)[r]{0}} \end{picture} \begin{picture}(1800,1511)(200,0) \special{" gnudict begin gsave 0 0 translate 0.100 0.100 scale 0 setgray newpath LTb 400 250 M 18 0 V 1332 0 R -18 0 V 400 520 M 18 0 V 1332 0 R -18 0 V 400 789 M 18 0 V 1332 0 R -18 0 V 400 1059 M 18 0 V 1332 0 R -18 0 V 400 250 M 0 18 V 0 993 R 0 -18 V 850 250 M 0 18 V 0 993 R 0 -18 V 1300 250 M 0 18 V 0 993 R 0 -18 V 1750 250 M 0 18 V 0 993 R 0 -18 V LTb 400 250 M 1350 0 V 0 1011 V -1350 0 V 400 250 L 1.000 UL LT0 1615 1126 M 106 0 V 29 -865 R -18 2 V -18 0 V -18 1 V -18 0 V -18 1 V -18 0 V -18 2 V -18 0 V -18 1 V -18 1 V -18 0 V -18 3 V -18 2 V -18 1 V -18 2 V -18 2 V -18 1 V -18 5 V -18 3 V -18 5 V -18 4 V -18 8 V -18 8 V -18 11 V -18 17 V -18 32 V -18 68 V -18 57 V -2 763 V -34 0 R 0 -11 V 1192 355 L -18 -22 V -18 -25 V -18 -16 V -18 -6 V -18 -5 V -18 -4 V -18 -4 V -18 -3 V -18 -1 V -18 -2 V -18 -2 V -18 -1 V -18 -1 V -18 -1 V -18 0 V -18 -2 V -18 0 V -18 -1 V -18 0 V -18 -1 V -18 0 V -18 0 V -18 -1 V -18 0 V -18 -1 V -18 0 V -18 0 V -18 0 V -18 -1 V -18 0 V -18 0 V -18 0 V -18 0 V -18 0 V -18 -1 V -18 0 V -18 0 V -18 0 V -18 0 V -18 0 V -18 0 V -18 0 V -18 -1 V -18 0 V 1.000 UL LT1 1615 1026 M 106 0 V 1681 261 M -10 0 V -11 0 V -10 2 V -10 -1 V -11 1 V -10 0 V -11 1 V -10 0 V -10 1 V -11 -1 V -10 1 V -11 0 V -10 1 V -11 0 V -10 1 V -10 0 V -11 0 V -10 1 V -11 2 V -10 -2 V -11 2 V -10 1 V -10 -1 V -11 1 V -10 2 V -11 0 V -10 2 V -10 0 V -11 1 V -10 0 V -11 2 V -10 2 V -11 -1 V -10 4 V -10 0 V -11 3 V -10 0 V -11 2 V -10 2 V -10 2 V -11 2 V -10 4 V -11 1 V -10 6 V -10 4 V -11 6 V -10 3 V -11 2 V -10 12 V -11 11 V -10 5 V -10 21 V -11 30 V -10 2 V -11 47 V -10 76 V -10 86 V -11 380 V -6 271 V -23 0 R -2 -590 V 1035 433 L -10 -40 V -10 -45 V -11 -18 V 994 312 L -11 -2 V -10 -9 V -11 -5 V -10 -6 V -10 -4 V -11 -2 V -10 -3 V -11 -2 V -10 -2 V -10 -2 V -11 -2 V -10 0 V -11 -2 V -10 -1 V -11 -1 V -10 -1 V -10 -1 V -11 0 V -10 -1 V -11 -1 V -10 -1 V -10 0 V -11 -1 V -10 0 V -11 -1 V -10 0 V -11 0 V -10 -1 V -10 0 V -11 -1 V -10 0 V -11 0 V -10 -1 V -10 0 V -11 0 V -10 0 V -11 -1 V -10 0 V -10 0 V -11 0 V -10 0 V -11 -1 V -10 0 V -11 0 V -10 0 V -10 0 V -11 -1 V -10 0 V -11 0 V -10 0 V -10 0 V -11 0 V -10 0 V -11 0 V -10 -1 V -11 0 V -10 0 V 1.000 UL LT2 1615 926 M 106 0 V 29 -667 R -5 0 V -7 0 V -6 0 V -7 1 V -6 0 V -7 0 V -7 0 V -6 1 V -7 0 V -6 -1 V -7 1 V -6 0 V -7 0 V -7 0 V -6 1 V -7 -1 V -6 1 V -7 1 V -6 -2 V -7 1 V -6 0 V -7 1 V -7 0 V -6 0 V -7 -1 V -6 2 V -7 -1 V -6 0 V -7 0 V -6 0 V -7 1 V -7 -1 V -6 2 V -7 0 V -6 -1 V -7 1 V -6 0 V -7 0 V -7 0 V -6 0 V -7 1 V -6 1 V -7 -1 V -6 1 V -7 0 V -6 0 V -7 1 V -7 0 V -6 1 V -7 0 V -6 0 V -7 0 V -6 1 V -7 0 V -6 1 V -7 0 V -7 1 V -6 0 V -7 0 V -6 1 V -7 0 V -6 0 V -7 2 V -7 -1 V -6 1 V -7 1 V -6 0 V -7 0 V -6 1 V -7 1 V -6 -1 V -7 3 V -7 0 V -6 0 V -7 1 V -6 2 V -7 0 V -6 0 V -7 2 V -6 1 V -7 0 V -7 1 V -6 2 V -7 1 V -6 1 V -7 1 V -6 3 V -7 0 V -7 2 V -6 2 V -7 2 V -6 2 V -7 3 V -6 0 V -7 3 V -6 3 V -7 2 V -7 2 V -6 4 V -7 4 V -6 3 V -7 4 V -6 5 V -7 4 V -6 5 V -7 8 V -7 10 V -6 11 V -7 9 V -6 8 V -7 14 V -6 21 V -7 12 V -7 16 V -6 32 V -7 50 V -6 68 V -7 66 V -6 202 V -7 389 V -18 0 R 944 940 L 938 721 L 931 537 L -6 -78 V -7 -41 V -6 -32 V -7 -19 V -6 -15 V -7 -9 V -7 -9 V -6 -12 V -7 -6 V -6 -5 V -7 -5 V -6 -2 V -7 -5 V -6 -3 V -7 -2 V -7 -3 V -6 -1 V -7 -3 V -6 -2 V -7 -1 V -6 -2 V -7 -1 V -7 -2 V -6 0 V -7 -2 V -6 -1 V -7 -1 V -6 0 V -7 -1 V -6 -2 V -7 0 V -7 -1 V -6 0 V -7 -1 V -6 -1 V -7 0 V -6 -1 V -7 0 V -6 -1 V -7 0 V -7 0 V -6 -1 V -7 0 V -6 -1 V -7 0 V -6 -1 V -7 0 V -7 0 V -6 -1 V -7 0 V -6 0 V -7 0 V -6 -1 V -7 0 V -6 0 V -7 0 V -7 -1 V -6 0 V -7 0 V -6 0 V -7 0 V -6 -1 V -7 0 V -6 0 V -7 0 V -7 0 V -6 -1 V -7 0 V -6 0 V -7 0 V -6 0 V -7 0 V -7 0 V -6 -1 V -7 0 V -6 0 V -7 0 V -6 0 V -7 0 V -6 0 V -7 -1 V 1.000 UL LT3 1615 826 M 106 0 V 1332 270 M -5 2 V -4 0 V -5 -1 V -4 2 V -4 1 V -5 1 V -4 -1 V -5 1 V -4 -1 V -4 2 V -5 -2 V -4 1 V -4 0 V -5 2 V -4 1 V -5 -2 V -4 1 V -4 1 V -5 -1 V -4 2 V -5 0 V -4 0 V -4 0 V -5 1 V -4 0 V -5 1 V -4 1 V -4 1 V -5 -1 V -4 0 V -5 3 V -4 0 V -4 -1 V -5 2 V -4 -1 V -5 1 V -4 1 V -4 0 V -5 2 V -4 -1 V -5 2 V -4 0 V -4 3 V -5 -3 V -4 2 V -5 3 V -4 0 V -4 1 V -5 -1 V -4 6 V -4 -3 V -5 1 V -4 1 V -5 4 V -4 -1 V -4 0 V -5 1 V -4 3 V -5 2 V -4 0 V -4 1 V -5 2 V -4 4 V -5 1 V -4 -2 V -4 5 V -5 1 V -4 1 V -5 6 V -4 -1 V -4 3 V -5 1 V -4 1 V -5 8 V -4 2 V -4 5 V -5 5 V -4 4 V -5 5 V -4 -2 V -4 6 V -5 5 V -4 4 V -5 15 V -4 11 V -4 -2 V -5 11 V -4 12 V -4 17 V -5 10 V -4 9 V -5 31 V -4 25 V -4 -5 V -5 39 V -4 45 V -5 31 V -4 31 V -4 181 V -5 35 V -4 300 V -3 91 V -22 0 R -2 -145 V 857 674 L -4 -64 V -5 -89 V -4 -66 V -5 -9 V -4 -22 V -4 -15 V -5 -25 V -4 -12 V -5 -10 V -4 -8 V -4 -8 V -5 -4 V -4 -8 V -4 -5 V -5 -5 V -4 -4 V -5 -1 V -4 -6 V -4 -2 V -5 -3 V -4 -2 V -5 -1 V -4 -3 V -4 -3 V -5 -1 V -4 -2 V -5 -2 V -4 -1 V -4 -2 V -5 -1 V -4 -1 V -5 -1 V -4 -2 V -4 0 V -5 -1 V -4 -2 V -5 0 V -4 -1 V -4 -1 V -5 -1 V -4 -1 V -5 0 V -4 -1 V -4 -1 V -5 0 V -4 -2 V -5 0 V -4 0 V -4 -1 V -5 0 V -4 0 V -4 -1 V -5 -1 V -4 -1 V -5 1 V -4 -1 V -4 -1 V -5 0 V -4 0 V -5 0 V -4 -1 V -4 -1 V -5 0 V -4 0 V -5 0 V -4 -1 V -4 0 V -5 0 V -4 0 V -5 0 V -4 -1 V -4 0 V -5 0 V -4 0 V -5 -1 V -4 0 V -4 0 V -5 -1 V -4 0 V -5 0 V -4 0 V -4 -1 V -5 0 V -4 0 V -5 0 V -4 0 V -4 0 V -5 0 V -4 -1 V -4 0 V -5 0 V -4 0 V -5 0 V -4 0 V -4 0 V -5 -1 V -4 0 V -5 0 V -4 0 V -4 0 V -5 0 V -4 0 V -5 -1 V -4 0 V 1.000 UL LT4 1615 726 M 106 0 V 1159 286 M -3 -3 V -3 -1 V -3 7 V -3 -3 V -3 2 V -3 -2 V -3 1 V -3 0 V -4 1 V -3 0 V -3 1 V -3 -1 V -3 1 V -3 2 V -3 -2 V -3 0 V -3 1 V -3 3 V -3 2 V -3 -2 V -4 1 V -3 -1 V -3 1 V -3 5 V -3 -3 V -3 0 V -3 2 V -3 1 V -3 -2 V -3 2 V -3 1 V -4 1 V -3 5 V -3 -6 V -3 1 V -3 2 V -3 3 V -3 0 V -3 -1 V -3 2 V -3 -2 V -3 2 V -3 7 V -4 -3 V -3 2 V -3 1 V -3 1 V -3 1 V -3 0 V -3 2 V -3 -3 V -3 7 V -3 1 V -3 -1 V -3 2 V -4 2 V -3 0 V -3 5 V -3 0 V -3 2 V -3 0 V -3 3 V -3 3 V -3 2 V -3 0 V -3 1 V -4 10 V -3 2 V -3 -6 V -3 6 V -3 4 V -3 1 V -3 -2 V -3 10 V -3 1 V -3 6 V -3 1 V -3 7 V -4 -1 V -3 7 V -3 7 V -3 9 V -3 4 V -3 1 V -3 4 V -3 15 V -3 1 V -3 13 V -3 7 V -4 8 V -3 14 V -3 8 V -3 22 V -3 1 V -3 5 V -3 24 V -3 58 V -3 15 V -3 -17 V -3 39 V -3 9 V -4 45 V -3 98 V -3 49 V -3 75 V -3 173 V -3 179 V 0 7 V -26 0 R 801 930 L 798 720 L -3 -85 V -3 -31 V -3 -37 V -3 -28 V -3 -54 V -3 -17 V -3 -22 V -4 -10 V -3 -25 V -3 -7 V -3 -14 V -3 -4 V -3 -3 V -3 -8 V -3 -11 V -3 -2 V -3 -10 V -3 -2 V -3 -6 V -4 -6 V -3 0 V -3 -4 V -3 -1 V -3 -5 V -3 -5 V -3 -2 V -3 -1 V -3 -3 V -3 -2 V -3 -2 V -4 -2 V -3 -1 V -3 -2 V -3 -3 V -3 0 V -3 -1 V -3 -3 V -3 -1 V -3 -1 V -3 0 V -3 -2 V -3 -2 V -4 -1 V -3 0 V -3 -1 V -3 -1 V -3 -1 V -3 -1 V -3 -1 V -3 -1 V -3 0 V -3 -1 V -3 -1 V -4 -1 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 -1 V -3 -1 V -3 -1 V -3 0 V -3 -1 V -3 0 V -3 -1 V -4 0 V -3 -1 V -3 0 V -3 0 V -3 -1 V -3 -1 V -3 0 V -3 0 V -3 0 V -3 -1 V -3 0 V -4 -1 V -3 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 -1 V -3 0 V -4 0 V -3 -1 V -3 0 V -3 0 V -3 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -4 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 0 V -3 0 V -3 0 V -3 -1 V -3 0 V -3 0 V -3 0 V -4 0 V -3 0 V -3 0 V -3 0 V -3 -1 V -3 0 V stroke grestore end showpage } \put(1565,726){\makebox(0,0)[r]{$\scriptstyle L=18$}} \put(1565,826){\makebox(0,0)[r]{$\scriptstyle L=16$}} \put(1565,926){\makebox(0,0)[r]{$\scriptstyle L=14$}} \put(1565,1026){\makebox(0,0)[r]{$\scriptstyle L=12$}} \put(1565,1126){\makebox(0,0)[r]{$\scriptstyle L=10$}} \put(1075,1411){\makebox(0,0){{\bf b)} $\quad \left[ \frac{\partial^2 s}{\partial m^2}\right]^{-1}({\scriptstyle \varepsilon,m_{\scriptscriptstyle {h=0}},L^{-1}})\,\,$ {\em vs.} $\varepsilon$\rule{7mm}{0mm}}} \put(1750,150){\makebox(0,0){-0.7}} \put(1300,150){\makebox(0,0){-0.8}} \put(850,150){\makebox(0,0){-0.9}} \put(400,150){\makebox(0,0){-1}} \put(350,1059){\makebox(0,0)[r]{1200}} \put(350,789){\makebox(0,0)[r]{800}} \put(350,520){\makebox(0,0)[r]{400}} \put(350,250){\makebox(0,0)[r]{0}} \end{picture} \begin{picture}(1800,1511)(300,0) \special{" gnudict begin gsave 0 0 translate 0.100 0.100 scale 0 setgray newpath LTb 350 250 M 18 0 V 1382 0 R -18 0 V 350 539 M 18 0 V 1382 0 R -18 0 V 350 828 M 18 0 V 1382 0 R -18 0 V 350 1117 M 18 0 V 1382 0 R -18 0 V 597 250 M 0 18 V 0 993 R 0 -18 V 926 250 M 0 18 V 0 993 R 0 -18 V 1256 250 M 0 18 V 0 993 R 0 -18 V 1585 250 M 0 18 V 0 993 R 0 -18 V LTb 350 250 M 1400 0 V 0 1011 V -1400 0 V 350 250 L 0.500 UP 1.000 UL LT0 1615 250 Pls 1455 564 Pls 1294 707 Pls 1133 809 Pls 972 899 Pls 811 976 Pls 650 1044 Pls 489 1108 Pls 827 770 Pls 0.500 UP 1.000 UL LT1 1713 250 Crs 1579 294 Crs 1445 570 Crs 1311 688 Crs 1177 781 Crs 1043 859 Crs 909 929 Crs 775 992 Crs 641 1048 Crs 507 1103 Crs 373 1152 Crs 827 670 Crs 0.500 UP 1.000 UL LT2 1733 250 Star 1624 250 Star 1515 471 Star 1405 605 Star 1296 697 Star 1186 775 Star 1077 840 Star 967 899 Star 858 954 Star 748 1003 Star 639 1049 Star 529 1092 Star 420 1133 Star 827 570 Star 0.500 UP 1.000 UL LT3 1646 250 Box 1538 437 Box 1431 583 Box 1323 678 Box 1215 757 Box 1107 824 Box 1000 883 Box 892 938 Box 784 986 Box 676 1033 Box 827 470 Box 0.500 UP 1.000 UL LT4 1727 250 BoxF 1615 250 BoxF 1503 497 BoxF 1391 619 BoxF 1279 712 BoxF 1167 788 BoxF 1055 853 BoxF 943 911 BoxF 831 965 BoxF 720 1013 BoxF 608 1058 BoxF 827 370 BoxF stroke grestore end showpage } \put(712,370){\makebox(0,0)[r]{$\scriptstyle L=18$}} \put(712,470){\makebox(0,0)[r]{$\scriptstyle L=16$}} \put(712,570){\makebox(0,0)[r]{$\scriptstyle L=14$}} \put(712,670){\makebox(0,0)[r]{$\scriptstyle L=12$}} \put(712,770){\makebox(0,0)[r]{$\scriptstyle L=10$}} \put(1050,1411){\makebox(0,0){{\bf c)} $\quad A({\scriptstyle L^{-1}})\,\,\,m_{\scriptscriptstyle {h=0}}({\scriptstyle \tilde{\varepsilon} ,L^{-1}})\,\,$ {\em vs.} $\tilde{\varepsilon} $\rule{10mm}{0mm}}} \put(1585,150){\makebox(0,0){0}} \put(1256,150){\makebox(0,0){-0.04}} \put(926,150){\makebox(0,0){-0.08}} \put(597,150){\makebox(0,0){-0.12}} \end{picture} \begin{picture}(1800,1511)(200,0) \special{" gnudict begin gsave 0 0 translate 0.100 0.100 scale 0 setgray newpath LTb 400 250 M 18 0 V 1332 0 R -18 0 V 400 452 M 18 0 V 1332 0 R -18 0 V 400 654 M 18 0 V 1332 0 R -18 0 V 400 857 M 18 0 V 1332 0 R -18 0 V 400 1059 M 18 0 V 1332 0 R -18 0 V 400 1261 M 18 0 V 1332 0 R -18 0 V 400 250 M 0 18 V 0 993 R 0 -18 V 738 250 M 0 18 V 0 993 R 0 -18 V 1075 250 M 0 18 V 0 993 R 0 -18 V 1413 250 M 0 18 V 0 993 R 0 -18 V 1750 250 M 0 18 V 0 993 R 0 -18 V LTb 400 250 M 1350 0 V 0 1011 V -1350 0 V 400 250 L 0.500 UP 1.000 UL LT0 1627 551 Pls 1495 486 Pls 1363 400 Pls 1232 328 Pls 1100 251 Pls 968 372 Pls 836 515 Pls 704 674 Pls 572 839 Pls 440 1019 Pls 1275 1160 Pls 0.500 UP 1.000 UL LT1 1729 642 Crs 1619 567 Crs 1509 486 Crs 1399 426 Crs 1289 369 Crs 1180 306 Crs 1070 252 Crs 960 384 Crs 850 498 Crs 740 620 Crs 630 756 Crs 521 905 Crs 411 1049 Crs 1275 1060 Crs 0.500 UP 1.000 UL LT2 1690 618 Star 1600 558 Star 1510 494 Star 1421 436 Star 1331 386 Star 1241 334 Star 1152 290 Star 1062 264 Star 972 353 Star 882 458 Star 793 560 Star 703 674 Star 613 787 Star 523 909 Star 434 1033 Star 1275 960 Star 0.500 UP 1.000 UL LT3 1740 645 Box 1652 596 Box 1563 538 Box 1475 457 Box 1386 420 Box 1298 367 Box 1210 321 Box 1121 276 Box 1033 289 Box 944 383 Box 856 480 Box 768 583 Box 679 694 Box 591 826 Box 502 928 Box 414 1069 Box 1275 860 Box 0.500 UP 1.000 UL LT4 1649 572 BoxF 1558 515 BoxF 1466 467 BoxF 1374 410 BoxF 1283 355 BoxF 1191 306 BoxF 1099 260 BoxF 1007 321 BoxF 916 413 BoxF 824 517 BoxF 732 634 BoxF 641 746 BoxF 549 882 BoxF 457 1016 BoxF 1275 760 BoxF stroke grestore end showpage } \put(1160,760){\makebox(0,0)[r]{$\scriptstyle L=18$}} \put(1160,860){\makebox(0,0)[r]{$\scriptstyle L=16$}} \put(1160,960){\makebox(0,0)[r]{$\scriptstyle L=14$}} \put(1160,1060){\makebox(0,0)[r]{$\scriptstyle L=12$}} \put(1160,1160){\makebox(0,0)[r]{$\scriptstyle L=10$}} \put(1075,1411){\makebox(0,0){{\bf d)} $\,\, B({\scriptstyle L^{-1}})\,\,\, \left[ \frac{\partial^2 s}{\partial m^2} \right] ({\scriptstyle \tilde{\varepsilon} ,m_{\scriptscriptstyle {h=0}},L^{-1}})\,\,$ {\em vs.} $\tilde{\varepsilon} $\rule{4mm}{0mm}}} \put(1750,150){\makebox(0,0){0.1}} \put(1413,150){\makebox(0,0){0.05}} \put(1075,150){\makebox(0,0){0}} \put(738,150){\makebox(0,0){-0.05}} \put(400,150){\makebox(0,0){-0.1}} \end{picture} {\caption[% ]{% \label{multib} \small{% Microcanonical magnetizations and "susceptibilities" of finite 3$d$-Ising systems (Figs.~a/b). For numerical convenience, $\left[ (\partial^2 s)/(\partial m^2)\right]^{-1}$ is shown instead of $\chi_{\scriptscriptstyle T;{h=0}}$. Note that both quantities show the same microcanonical finite-size scaling behaviour and -- in the thermodynamic limit -- the same critical behaviour. Multiplication of $m_{\scriptscriptstyle {h=0}}$ and $(\partial^2 s)/(\partial m^2)$ by suitable scaling factors $A({\scriptstyle L^{-1}})$ and $B({\scriptstyle L^{-1}})$ and plotting the thus obtained results against $\tilde{\varepsilon}$ yields data collapse (Figs.~c/d). Whereas Figs.~a/b suggest the introduction of finite system exponents, Figs.~c/d give rise to the assumption that the thus defined exponents show no system size dependence. } }} \end{figure} \end{center} \vspace{-8mm} Data collapse is achieved by suitable scaling of the amplitudes of $m_{\scriptscriptstyle {h=0}}$ and $\chi_{\scriptscriptstyle T;{h=0}}$. Therefore, it seems to be evident that the finite system critical exponents are independent of the system size (i.e.~take on the same value $\forall L^{-1} \ne 0$), but differ from the values expected in the thermodynamic limit ({\bf TDL}), $\beta_{\varepsilon,\infty} \approx 0.37$, $\gamma_{\varepsilon,\infty}\approx 1.38$, whereas from numerical data we obtain $\beta_{\varepsilon,L}\approx 0.5$, $\gamma_{\varepsilon,L}\approx 1$ for all $L$ considered. \noindent In this paper, a microcanonical finite size scaling ({\bf MFSS}) theory is developed, taking into account the following constraints (partly justified above):\\ \vspace{-.6cm} \newcounter{con} \begin{list}{(\roman{con})}{ \usecounter{con} \setlength{\leftmargin}{.8cm} \setlength{\rightmargin}{0cm} \setlength{\labelsep}{0.25cm} \setlength{\itemindent}{0cm} \setlength{\topsep}{2ex} } \item consistence with the canonical finite size scaling ({\bf CFSS}) theory\\[-2ex \item power law behaviour of the finite system magnetization and susceptibility\\[-2ex \item finite system critical exponents which do not depend on the system size and which are not necessarily identical to those of the infinite system\\[-2ex \end{list} \section{Microcanonical finite-size scaling (MFSS)} In the vicinity of a critical point ($t, h, \varepsilon^*, m \approx 0$), the thermodynamic potentials can be divided into a singular part (superscript$\phantom{a}^*$) which describes the non-analytic behaviour, and a regular part (subscript $\phantom{a}_{reg}$). For the free energy density and the specific entropy this reads: \begin{equation} \label{free_energy} \hspace{-5mm} g(t,h) \, =\, g^*(t,h) \,+\, g_{reg}(t,h) \quad , \quad {\scriptstyle t:=\frac{T-T_c}{T_c} } \quad , \quad \begin{array}{l} {\scriptstyle T_c}\, \mbox{\scriptsize critical temperature} \\ {\scriptstyle h}\,\, \mbox{\scriptsize magnetic field} \end{array} \end{equation} \begin{equation} \label{entropy} \hspace{-5mm} s(\varepsilon^*,m) \,=\, s^*(\varepsilon^*,m) \,+\, s_{reg}(\varepsilon^*,m) \quad , \quad {\scriptstyle \varepsilon^* := \frac{\varepsilon-\varepsilon_c}{|\varepsilon_c|} } \quad , \quad {\scriptstyle \varepsilon_c} \,\,\mbox{\scriptsize critical enthalpy} \end{equation} It can be shown (see e.g.~\cite{CJL76} and references therein) that in the TDL the singular parts of the various potentials are homogeneous functions (cf.~Eqs.~(6),(7) in Fig.~2) and all critical exponents can be expressed in terms of the degrees of homogeneity ($a_\varepsilon=(1-\alpha)/(2-\alpha)$, $a_m=1/(\delta+1)$, $a_t=1-a_\varepsilon$, $a_h=1-a_m$). \setcounter{equation}{9} The equivalence of ensembles is valid only in the TDL. Hence, the thermodynamic potentials of finite systems have to be classified as \underline{c}anonical or \underline{m}i\-cro\-ca\-no\-ni\-cal quantities (cf.~\cite{KPH98},\cite{KPHpub}). Starting point for the CFSS is the so-called ``CFSS assumption'' \cite{FB72}: For large but finite systems, $g_c^*(t,h,L^{-1})$ is a homogeneous function in accordance with Eq.~(8). \\ We have shown elsewhere \cite{KPHpub} that the MFSS assumption (9) entails (8), whereas proof of the reverse is more difficult. Note that Eq.~(9) even accounts for the possibility of $s_m^*$ showing no system size dependence at all; nevertheless, this case results in an $L$-dependence of $g_c^*$ and in CFSS. In the rest of this paper, we will discuss the consequences of the MFSS assumption which states that: $$ \hspace{-10mm} \mbox{\bf MFSS assumption}: \qquad \begin{array}{c} s_m^*({\scriptstyle \varepsilon^*,m,L^{-1}})\mbox{ is a homogeneous function}\\[1ex] \mbox{of its arguments (cf.~Fig.~2, Eq.~(9))} \end{array} $$ \def\scriptstyle{\scriptstyle} \def\scriptscriptstyle{\scriptscriptstyle} \begin{figure}[ht] \begin{center} \unitlength1cm \begin{picture}(15,7)(2.3,0) \put(1.8,5.25){\framebox(4,.7){}} \put(1.8,3.25){\framebox(4,.7){}} \put(1.8,.3){\framebox(4,.7){}} \put(10,5.25){\framebox(4.6,.7){}} \put(10,3.25){\framebox(4.6,.7){}} \put(10,.3){\framebox(4.6,.7){}} \thicklines \put(1.8,6.3){$ \scriptstyle (\mathbf{6}) \,\, g^*(t,h)\,\,=\,\,\lambda^{-1} g^*(\lambda^{a_t}t,\lambda^{a_h}h) $} \put(2.5,5.5){Homogeneity of $g^*$} \put(3.7,4){\vector(0,1){1.2}} \put(2.7,4.5){$\scriptscriptstyle L \to \, \infty$} \put(2.5,3.5){CFSS assumption} \put(1.8,2.8){$\scriptstyle (\mathbf{8}) \qquad g_c^*(t,h,L^{-1})\,\,=$} \put(1.8,2.3){$\scriptstyle \,\,\quad \lambda^{-1} g_c^*(\lambda^{a_t}t,\lambda^{a_h}h,\lambda^{1/d}L^{-1})$} \put(10,6.3){$\scriptstyle (\mathbf{7})\,\, s^*(\varepsilon^*,m)\,\,=\,\, \lambda^{-1} s^*(\lambda^{a_\varepsilon}\varepsilon^*, \lambda^{a_m}m)$} \put(10.8,5.5){Homogeneity of $s^*$} \put(12.2,4){\vector(0,1){1.2}} \put(12.5,4.5){$\scriptscriptstyle L \to \, \infty$} \put(10.8,3.5){MFSS assumption} \put(10,2.8){$\scriptstyle (\mathbf{9}) \qquad s_m^*(\varepsilon^*,m,L^{-1})\,\,\,=$} \put(10,2.3){$\scriptstyle \,\,\quad \lambda^{-1} s_m^*(\lambda^{a_\varepsilon}\varepsilon^*, \lambda^{a_m}m, \lambda^{1/d}L^{-1})$} \put(12.2,1.9){\vector(0,-1){.8}} \put(10.2,.6){\scriptsize \underline{M}icrocanonical \underline{F}inite \underline{S}ize \underline{S}caling } \put(3.7,1.9){\vector(0,-1){.8}} \put(2,.6){\scriptsize \underline{C}anonical \underline{F}inite \underline{S}ize \underline{S}caling } \put(7,3.8){$\scriptstyle {via~Laplace~T.}$} \put(9.5,3.6){\vector(-1,0){3}} \put(7,5.8){$\scriptstyle {via~Legendre~T.}$} \put(6.5,5.6){\vector(1,0){3}} \put(9.5,5.6){\vector(-1,0){3}} \end{picture} \vspace{-5mm} {\caption[% ]{% \label{fig1} \small{% Homogeneity relations $(\mbox{valid}\; \forall \lambda\, \! > 0)$ for the singular parts of the free energy and the entropy for finite (Eqs.~(8),(9)) as well as infinite systems (Eqs.~(6),(7)). In the TDL, a Legendre transform connects the homogeneity of $g^*$ to the homogeneity of $s^*$ (see Ref.~\cite{S71}), whereas for finite systems it can be shown that the CFSS assumption is a consequence of the MFSS assumption, i.e., MFSS is consistent with CFSS \cite{KPHpub}. } }} \end{center} \end{figure} \noindent From the MFSS assumption (9) and Eqs.~(\ref{mmic}) and (\ref{chimic}), the {\it MFSS relations of the magnetization and the susceptibility} are derived easily \cite{FN1}: \begin{equation} \label{mfss} m^*_{\scriptscriptstyle {h=0}} ({\scriptstyle \varepsilon^*,L^{-1}} ) \,\, = \,\, \lambda^{-a_m}m^*_{\scriptscriptstyle {h=0}} ({\scriptstyle \lambda^{a_\varepsilon}\varepsilon^* ,\lambda^{1/d}L^{-1}} ) \,\,\stackrel{\lambda = L^d}{=} \,\, L^{-da_m} \Phi_{m^*} ({\scriptstyle L^{da_\varepsilon}\varepsilon^*}) \end{equation} \begin{equation} \label{chifss} \hspace{-6mm} \chi_{\scriptscriptstyle T;{h=0}}^*({\scriptstyle \varepsilon^* , L^{-1} } ) \, = \, \lambda^{1-2a_m} \chi_{\scriptscriptstyle T;{h=0}}^*({\scriptstyle \lambda^{a_\varepsilon}\varepsilon^* , \lambda^{1/d}L^{-1} } ) \, \stackrel{\lambda = L^d}{=} \, L^{d(1-2a_m)} \Phi_{\chi^*} ({\scriptstyle L^{da_\varepsilon}\varepsilon^*} ) \end{equation} \noindent where the $\Phi_i$ are so-called MFSS functions which describe the behaviour of the magnetization and the susceptibility of finite systems in the vicinity of the critical point $\varepsilon_c$ of the infinite system. \noindent {\it MFSS of the transition point:} From Figs.~1a/c, it can be deduced that the MFSS function of the magnetization $\Phi_m(x)$ is zero for $L^{da_\varepsilon}\varepsilon^* =: x \ge x_T$, where $x_T$ determines the magnetic transition point $\varepsilon_T(L)$ for finite systems: \begin{equation} \label{efss} \hspace{-5mm} x_T\,=\,L^{da_\varepsilon} \frac{\varepsilon_T(L)-\varepsilon_c}{|\varepsilon_c|} \,\,\,\, \Longleftrightarrow \,\,\,\, \varepsilon_T(L) \,=\, \varepsilon_c (1+\tilde{x}_TL^{-da_\varepsilon}) \, , \,\, \tilde{x}_T:=\frac{\varepsilon_c}{|\varepsilon_c|}x_T \end{equation} \noindent\underline{Remarks:} a) An additional quantity $\tilde{s}_m^*$ can be introduced, which is defined to be the singular part of the entropy, written in terms of the reduced enthalpy $\tilde{\varepsilon}(L) := (\varepsilon-\varepsilon_T(L))/|\varepsilon_T(L)|$. \newpage \noindent Using Eq.~(\ref{efss}), the thus defined entropy $\tilde{s}_m^*$ can be shown to possess the same degrees of homogeneity as $s_m^*$: \begin{equation} \label{stild} \tilde{s}_m^*({\scriptstyle \tilde{\varepsilon},m,L^{-1}} ) \,=\,\lambda^{-1}\tilde{s}_m^*({\scriptstyle \lambda^{a_\varepsilon}\tilde{\varepsilon},\lambda^{a_m}m,\lambda^{1/d}L^{-1}} ) \,\,\, , \,\,\,\mbox{where} \end{equation} $$ \tilde{s}_m^*({\scriptstyle \tilde{\varepsilon},m,L^{-1}} ) \,\,:=\,\,s_m^*({\scriptstyle \varepsilon^*=\tilde{\varepsilon}+x_TL^{-da_\varepsilon},m,L^{-1}})\,\,\, . \phantom{..........................................} $$ Starting from expression (\ref{stild}), MFSS relations can be derived for the magnetization and the susceptibility. They illustrate the system size dependence of the behaviour of $m_{\scriptscriptstyle {h=0}}$ and $\chi_{\scriptscriptstyle T;{h=0}}$ in the vicinity of the transition point $\varepsilon_T(L)$ of the finite system: \begin{equation} \label{mfsstild} \tilde{m}_{\scriptscriptstyle {h=0}} ({\scriptstyle \tilde{\varepsilon},L^{-1}} ) \,\, =\,\, \lambda^{-a_m}\tilde{m}_{\scriptscriptstyle {h=0}} ({\scriptstyle \lambda^{a_\varepsilon} \tilde{\varepsilon} ,\lambda^{1/d}L^{-1}} ) \,\, \stackrel{\lambda = L^d}{=} \,\, L^{-da_m} \Phi_{\tilde{m}} ({\scriptstyle L^{da_\varepsilon}\tilde{\varepsilon}} ) \end{equation} \begin{equation} \label{chifsstild} \hspace{-5mm} \tilde{\chi}_{\scriptscriptstyle T;{h=0}}({\scriptstyle \tilde{\varepsilon} , L^{-1} } ) \,\, = \,\, \lambda^{1-2a_m} \tilde{\chi}_{\scriptscriptstyle T;{h=0}}({\scriptstyle \lambda^{a_\varepsilon}\tilde{\varepsilon} , \lambda^{1/d}L^{-1} } ) \,\, \stackrel{\lambda = L^d}{=} \,\, L^{d(1-2a_m)} \Phi_{\tilde{\chi}} ({\scriptstyle L^{da_\varepsilon}\tilde{\varepsilon}} ) \end{equation} b) Numerical data suggest power law behaviour for both the magnetization and the susceptibility of finite systems. Therefore, the MFSS functions $\Phi_{\tilde{m}}$ and $\Phi_{\tilde{\chi}}$ have to be power laws, governed by the respective finite system exponents which are \underline{not} determined by the degrees of homogeneity of $s_m^*$: \begin{eqnarray} \Phi_{\tilde{m}} (x) & \propto & (-x)^{\beta_{\varepsilon,L}} \;\;\;\;\;\;\;\;\mbox{for}\;\;x\,\le\,0 \\ \Phi_{\tilde{\chi}} (x) & \propto & |x|^{-\gamma_{\varepsilon,L}} \end{eqnarray} c) $g_c^*$ is analytic for all finite $L$ and shows non-analyticities only in the TDL. To the best of our knowledge, no proof exists that $s_m^*$ is analytic for finite systems. Depending on the values of $\beta_{\varepsilon,L}$ and $\gamma_{\varepsilon,L}$, Eqs.~(16) and (17) imply the possibility for $s_m^*$ to be either an analytic or a non-analytic function. This means that even for a completely analytic entropy, non-analyticities can occur for the microcanonical magnetization and susceptibility.\\ d) Note that it is indeed possible to explicitly construct a function $\tilde{s}_m^*$ which complies with the requirements (i)-(iii) stated at the end of the introduction. This explicit form is quite informative with respect to the "sudden" change of the exponents from their finite system values towards their values in the TDL, but has to be discussed elsewhere \cite{KPHpub}. \section{Conclusion} A MFSS theory has been developed in accordance with the demanded properties (i)-(iii) stated above. Amazingly, although the scaling laws (7) and (9) comprise identical degrees of homogeneity $a_\varepsilon$ and $a_m$ for the singular parts of the entropy of the finite \underline{and} infinite system respectively, they nevertheless can account for power law behaviour with different exponents $\beta_{\varepsilon,L}$, $\gamma_{\varepsilon,L}$ for finite and $\beta_{\varepsilon,\infty}$, $\gamma_{\varepsilon,\infty}$ for infinite systems. As canonical potentials emerge from the microcanonical ones basically by means of a Laplace transform, it is to be expected that the smoothing properties of this integral transform cause MFSS to be applicable for smaller systems than CFSS.
\section{Introduction} The particles in most granular materials carry a net electrical charge. This charge emerges naturally due to contact electrification during transport or is artificially induced in industrial processes. It is well known~\cite{Kanazawa,Nieh88,Singh85}, for instance, that particles always charge when transported through a pipe. In industry, contact electrification is used for dry separation of different plastic materials or salts~\cite{Inculet}, which tend to get oppositely charged and hence are deflected into opposite directions when falling through a condenser. Another application is powder varnishing, where uniformly charged pigment particles are blown towards the object to be painted, which is oppositely charged. Whereas the dynamics of electrically neutral grains have been studied in great detail, little is known about what will change, if the grains are charged. In this paper we present the answer for collisional cooling, a basic phenomenon, which is responsible for many of the remarkable properties of dilute granular media. By collisional cooling one means that the relative motion of the grains, which lets them collide and can be compared to the thermal motion of molecules in a gas, becomes weaker with every collision, because energy is irreversibly transferred to the internal degrees of freedom of the grains. In 1983 Haff~\cite{Haff} showed, that the rate, at which the kinetic energy of the relative motion of the grains is dissipated in a homogeneous granular gas, is proportional to $T^{3/2}$, where $T$ is the so called granular temperature. It is defined as the mean square fluctuation of the grain velocities divided by the space dimension: \begin{equation} T = \langle \vec v\,^2 - \langle \vec v \rangle^2 \rangle/3 . \label{eq1} \end{equation} A consequence of this dissipation rate is that the granular temperature of a freely cooling granular gas decays with time as $t^{-2}$. We shall discuss, how these laws change, if the particles are uniformly charged. Due to the irreversible particle interactions large scale patterns form in granular media, such as planetary rings \cite{Planets} or the cellular patterns in vertically vibrated granular layers \cite{Umbanhowar}. This happens even without external driving \cite{Goldhirsch,McNamara,Luding99}, where one can distinguish a kinetic, a shearing and a clustering regime. The regimes depend on the density, the system size and on the restitution coefficient $e_{\rm n}=-v_{\rm n}'/v_{\rm n}$, which is the ratio of the normal components of the relative velocities before and after a collision between two spherical grains. The $T^{3/2}$ cooling law holds, provided the restitution coefficient may be regarded as independent of $v_{\rm n}$ \cite{Brilliantov}, and the system remains approximately homogeneous \cite{Luding98}. The latter condition defines the kinetic regime, which is observed for the highest values of the restitution coefficient, whereas the two other regimes are more complicated because of the inhomogeneities. Such inhomogeneities can only occur as transients, if all particles are equally charged, because the Coulomb repulsion will homogenise the system again. In order to avoid additional dissipation mechanisms due to eddy currents within the grains we consider only insulating materials. Unfortunately, up to now, no consistent microscopic theory for contact electrification of insulators exists~\cite{Lowell}. In powder processing two types of charge distribution are observed~\cite{Singh85}: A bipolar charging, where the charges of the particles in the powder can have opposite sign and the whole powder is almost neutral. The other case is monopolar charging, for which the particles tend to carry charges of the same sign and the countercharge is transferred to the container walls. It depends largely on the type of processing, whether one observes bipolar or monopolar charging, which means, that the material of the container, the material of the powder and other more ambiguous things, like air humidity or room temperature are important~\cite{Lowell}. The outline of this article is as follows: The next section specifies the model we are considering. A simple derivation of the dissipation rate in dilute charged granular media based on kinetic gas theory is given in section~\ref{sec_ANALY}. We find, that the dissipation rate is essentially the one known from uncharged granular media multiplied with a Boltzmann factor. Section~\ref{sec_DELTA} compares the analytic results with computer simulations. We find that in non-dilute systems the Coulomb repulsion is effectively reduced. This reduction will be explained, and we determine its dependence on the solid fraction of the granular gas. In the appendix we discuss the new simulation method we developed for this investigation. It is a molecular dynamics method, that avoids the so called {\em brake-failure\/}~\cite{Schaefer}. \section{The model} In this paper, we consider monopolar charging, which is the usual case if insulators are transported through a metal pipe~\cite{Nieh88,Singh85}. For simplicity we assume, that all particles have the same point charge $q$ centred in a sphere of diameter $d$ and mass $m$. No polarisation and no charge transfer during contact will be considered. The particle velocities are assumed to be much smaller than the velocity of light, so that relativistic effects (retardation and magnetic fields due to the particle motion) can be neglected. The electrodynamic interaction between the particles can then be approximated by the Coulomb potential: \begin{equation}\label{Coulomb} \Phi _{ij} = q^2 / r_{ij}, \end{equation} where $r_{ij}$ is the distance between the centers of particles $i$ and $j$. We consider the collisions as being instantaneous, which is a good approximation for the dilute granular gas, where the time between collisions is much longer than the duration of the contact between two particles. As the incomplete restitution ($e_{\rm n}<1$) is the main dissipation mechanism in granular gases, Coulomb friction will be neglected in this paper. Also, the dependence of the restitution coefficient on the relative velocity \cite{Schaefer2,Brilliantov} will be ignored, so that the constant $e_{\rm n}$ is the only material parameter in our model. The particles are confined to a volume $V = L^3$ with periodic boundary conditions in all three directions. The periodic volume can be thought of as a sufficiently homogeneous subpart of a larger system, which is kept from expanding by reflecting walls. For vanishing particle diameter this model corresponds to the One Component Plasma (OCP)~\cite{Baus}. In the OCP a classical plasma is modelled by positive point charges (the ions) acting via the Coulomb potential, whereas the electrons are considered to be homogeneously smeared out over the whole system. In the OCP the electron background cannot be polarised, i.e. Debye screening does not exist, as is the case in our model, too. \section{Analytical results for dilute systems} \label{sec_ANALY} In this section we derive an approximate expression for the dissipation rate in a dilute system of equally charged granular spheres, neglecting particle correlations. Basically we apply the kinetic gas theory, but include inelastic collisions. Using the analytic form of the dissipation rate in the dilute limit derived here, we will discuss the dissipation in a non-dilute system, where correlations are important, in the next chapter. We start with calculating the collision frequency of a fixed particle $i$ with any of the other particles $j$. If they were not charged, two particles would collide provided the relative velocity $\vec u $ points into the direction of the distance vector $\vec r = \vec r_j - \vec r_i$ connecting the particle centers, $\vec u \cdot \vec r > 0$, and the impact parameter $b = |\vec r \times \vec u|/u $ is smaller than the sum of the particle radii, $b \leq b_{\rm max}=d$. If the particles carry the charge $q$, they repel each other and the maximum impact parameter $b_{\rm max}$ becomes smaller than $d$ (see Fig.\ref{fig1}). By the conservation laws for angular momentum and for energy one gets: \begin{equation} \label{bmax} b_{\rm max}^2 = d^2 \left( 1 - \frac{2 E_{\rm q}}{\mu \, u^2} \right) \end{equation} where $\mu = m/2$ is the reduced mass. $E_{\rm q} = q^2/d$ denotes the energy barrier which must be overcome to let two particles collide in the dilute limit. It is the difference of the potential energies at contact and when they are infinitely far apart. Eq.~(\ref{bmax}) is independent of the actual form of the potential, as long as it has radial symmetry. (Note that energy is conserved as long as the particles do not touch each other.) \begin{figure}[tb] \centerline{\psfig{figure=FIG1.ps,angle=270,width=8cm}} \caption{Particle $i$ collides with particle $j$.} \label{fig1} \end{figure} Imagine a beam of particles, all having the same asymptotic velocity $\vec u $ far away from particle $j$. All particles within an asymptotic cylinder of radius $b_{\rm max}$ around the axis through the center of $j$ with the direction of $\vec u $ will collide with particle $j$. There will be $\pi \, b_{\rm max}^2 \, u \, n$ such collisions per unit time, where $n=N/V$ is the number density. Integrating over all relative velocities $\vec u$ gives the collision frequency of a single particle in the granular gas in mean field approximation: \begin{equation} f = \pi n \int \limits_{u \geq u_0} d^3\!u \, u \, b_{\rm max}^2(u) \, p(u ). \end{equation} $u_0 = \sqrt{2 E_{\rm q}/\mu}$ is the minimal relative velocity at infinity, for which a collision can occur overcoming the repulsive interaction. We assume that the particle velocity distribution is Gaussian with variance $3 T$ (see (\ref{eq1})), so that the relative velocity will have a Gaussian distribution as well, with \begin{equation} \langle u^2 \rangle= 6 \, T \end{equation} Hence, the total number of binary collisions per unit time and per unit volume is given by: \begin{equation} \label{Ng} \dot N_g = 1/2\, f\, n = 2\sqrt{\pi} \, n^2 \, d^2 \, \sqrt{T} \cdot \exp\left(- \frac{E_{\rm q}}{mT} \right) \end{equation} The factor $1/2$ avoids double counting of collisions. This corresponds to textbook physics for chemical reaction rates as can be found for example in Present\cite{Present}. Now we calculate the dissipation rate: The energy loss due to a single inelastic collision is: \begin{equation} \delta \!E(u, b) = \frac{\mu}{2} \, \left(1-e_{\rm n}^2\right) \, {u_{\rm n}^*}^2 \end{equation} where $u_{\rm n}^{*}$ means the normal component of the relative velocity $\vec u^{*}$ at the collision. It can be calculated easily from ${u_{\rm n}^{*}}^2={u^{*}}^2 - {u_{\rm t}^{*}}^2$: The tangential component is determined by angular momentum conservation, \begin{equation} \mu u b = \mu u_{\rm t}^{*} d, \end{equation} and energy conservation gives \begin{equation} {u^*}^2 = u^2 \left( 1 - \frac{2 E_{\rm q}}{\mu \, u^2} \right) . \end{equation} This yields \begin{equation} {u_{\rm n}^{*}}^2 = u^2 \left( 1 - \frac{b^2}{d^2} - \frac{2 E_{\rm q}}{\mu \, u^2} \right) \end{equation} The energy loss in one collision is therefore: \begin{equation} \delta \!E (u,b) = \frac{\mu}{2} \, \left(1-e_{\rm n}^2\right) \, u^2 \, \left( 1 - \frac{b^2}{d^2} - \frac{2 E_{\rm q}}{\mu \, u^2} \right) \end{equation} Assuming a homogeneous distribution of particles, we eliminate the $b$-dependence by averaging over the area $\pi b_{\rm max}^2$ (see Fig.\ref{fig1}): \begin{eqnarray} \delta \!E (u) & = & \frac{1}{\pi b_{\rm max}^2} \int \limits_{0}^{b_{\rm max}} db \, 2\pi b \, \delta \!E (u,b) \\ & = & \frac{\mu}{4} \, u^2 \, \left(1-e_{\rm n}^2\right) \left( 1 - \frac{2 E_{\rm q}}{\mu \, u^2} \right) \label{eq11b} \end{eqnarray} The dissipated energy per unit time due to collisions with relative velocity $u$ is then the number of such collisions per unit volume, ${1}/{2} \, n^2 \pi b_{\rm max}^2 \, u$, times the energy loss $\delta \!E$, Eq.~(\ref{eq11b}). Finally we get the dissipation rate per unit volume in the dilute limit ($\nu \to 0$) by integration over the relative velocity distribution: \begin{eqnarray} \gamma & = & \frac{\pi}{2} n^2 \int \limits_{u \geq u_0} d^3u \, b_{\rm max}^2 \, u \, \delta \!E (u) \, p(u) \nonumber \\ \label{gamma_tot} & = & 2\sqrt{\pi} \, n^2 d^2 m \, \left(1-e_{\rm n}^2\right) \, T^{3/2} \cdot \exp\left(- \frac{E_{\rm q}}{m T}\right) \end{eqnarray} The dissipation rate of an {\em uncharged\/} granular system in the dilute limit in the kinetic regime is given by~\cite{Haff}: \begin{equation} \label{gamma_0} \gamma_0 = 2\sqrt{\pi} \, n^2 d^2 m \, \left(1-e_{\rm n}^2\right) \, T^{3/2} \end{equation} Thus the dissipation rate (\ref{gamma_tot}) in a monopolar charged dilute granular gas and the one for the uncharged case differ only by a Boltzmann factor, $\gamma = \gamma_0 \cdot \exp\left(- {E_{\rm q}}/{m T}\right)$. This is the main result of the analytic treatment in this section. It remains valid for any repulsive pair interaction between the grains that has rotational symmetry. \section{Dissipation rate for dense systems} \label{sec_DELTA} In order to discuss the dissipation rate $\gamma$ in a non-dilute system of charged granular matter, let us recall the analytic form of $\gamma$ in an uncharged non-dilute system. The derivation is basically done by using the Enskog expansion of the velocity distribution function for dense gases~\cite{Lun84}. One gets the dissipation rate for a non dilute uncharged system: \begin{equation} \label{gamma_unch.dns} \gamma = \gamma_0 \cdot g_{\rm hs}(\nu) \end{equation} where $\gamma_0$ is given by Eq.~(\ref{gamma_0}) and $g_{\rm hs}(\nu)$ is the equilibrium pair distribution function of the non-dissipative hard-sphere fluid at contact. It only depends on the solid fraction $\nu = \pi n d^3/ 6$: \begin{equation} g_{\rm hs}(\nu) = \frac{2-\nu}{2 (1-\nu)^3} \end{equation} (Carnahan and Starling~\cite{Carnahan}, Jenkins and Richman \cite{Jenkins}). Our system consists of dissipative charged hard-spheres (CHS). The Boltzmann factor in Eq.~(\ref{gamma_tot}) is just the equilibrium pair distribution function at contact in the dilute limit for a CHS-fluid, $\lim_{\nu \to 0} g_{\rm chs}(\nu,q) = \exp \left(- E_{\rm q}/m T\right)$. So it is plausible, that the dissipation rate for a dense system of dissipative CHS is: \begin{equation} \label{gamma_ch.dns} \gamma = \gamma_0 \cdot g_{\rm chs}(\nu, q) \end{equation} Unfortunately the literature is lacking a satisfying analytic expression for $g_{\rm chs}$. In 1972 Palmer and Weeks~\cite{Palmer72} did a mean spherical model for the CHS and derived an analytic expression for $g_{\rm chs}$, but this approximation is poor for low densities. Many methods~\cite{bunch_of_chs} give $g_{\rm chs}$ as a result of integral equations, that can be solved numerically. We do not use those approximations, but make the following ansatz for $g_{\rm chs}$: \begin{equation} \label{gamma_ansatz} g_{\rm chs}(\nu,q) \approx g_{\rm hs}(\nu) \cdot \exp\left(- \frac{E_{\rm eff}(\nu)}{m T}\right) \end{equation} As in the dilute case we assume that the long range Coulomb repulsion modifies the pair correlation function of the uncharged hard sphere gas by a Boltzmann factor. Note that the granular temperature enters the pair correlation function only through this Boltzmann factor. The hard core repulsion is not connected with any energy scale, so that the pair correlation function $g_{\rm hs}$ cannot depend on $T$. The effective energy barrier $E_{\rm eff}$ must approach $E_{\rm q}$ in the dilute limit. Hence the ansatz (\ref{gamma_ansatz}) contains both the uncharged and the dilute limit, (\ref{gamma_unch.dns}) respectively (\ref{gamma_tot}). In order to check the ansatz (\ref{gamma_ansatz}) we did computer simulations using the MD algorithm as described in the appendix. Test systems of varying solid fraction $\nu$ and particle number ranging from $N=256$ to $N=1024$ were prepared at a starting temperature $T_0$. As soon as the simulation starts, the granular temperature drops because of the inelastic collisions. We measured the dissipation rate $\gamma$ and the granular temperature during this evolution. According to Eq.~(\ref{gamma_ansatz}) and Eq.~(\ref{gamma_ch.dns}) the dissipation rate is $\gamma = \gamma_0 \, g_{\rm hs}(\nu) \cdot \exp(E_{\rm eff}(\nu) /mT)$. An Arrhenius plot ($\ln(\gamma/\gamma_0\, g_{\rm hs})$ versus $E_{\rm q}/mT$) should give a straight line whose negative slope is the effective energy barrier $E_{\rm eff}$. Fig.~\ref{work.fig.2} shows two examples of these simulations. The Arrhenius plots are linear to a very good approximation. This confirms the ansatz (\ref{gamma_ansatz}). Systems with high densities show slight deviations from linearity. \begin{figure}[tb] \centerline{\psfig{figure=FIG2.ps,angle=270,width=8cm}} \caption{Arrhenius-plot of the dissipation rate $\gamma$ normalised by the one of the uncharged system, Eq.~(\ref{gamma_unch.dns}). Granular temperature is scaled by $E_{\rm q}/m$. Filled circles correspond to simulations of density $\nu = 3.375 \cdot 10^{-3}$ and filled squares $\nu = 7 \cdot 10^{-2}$. The linear fits yield: $E_{\rm eff}/E_{\rm q} = 0.70$ for the lower density and $E_{\rm eff}/E_{\rm q} = 0.27$ in the other case.} \label{work.fig.2} \end{figure} The negative slopes $E_{\rm eff}/E_{\rm q}$ in Fig.~\ref{work.fig.2} are smaller than $1$, which means, that the effective energy barrier is smaller than in the dilute system. The explanation is that two particles which are about to collide not only repel each other but are also pushed together by being repelled from all the other charged particles in the system. For dimensional reasons the effective energy barrier to be overcome, when two particles collide, must be of the form \begin{equation} \label{Phi} E_{\rm eff} = \frac{q^2}{d} - \frac{q^2}{\ell} f(d/\ell), \end{equation} where $\ell>d$ is the typical distance between the charged particles and $f$ is a dimensionless function. The first term is the Coulomb interaction $E_{\rm q}$ of the collision partners at contact. The second term takes the interaction with all other particles in the system into account. It is negative, because the energy barrier for the collision is reduced in dense systems. Obviously, for a dense packing, $\ell \rightarrow d$, the energy barrier for a collision must vanish, i.e. $ E_{\rm eff}|_{d=\ell} = 0 $. Moreover, if one takes a dense packing and reduces the radii of all particles infinitesimally, keeping their centers in place, all particles should be force free for symmetry reasons. Therefore, the energy barrier must vanish at least quadratically in $(\ell - d)$, i.e. $\partial E_{\rm eff}/\partial d |_{d=\ell} =0$. For the function $f$ this implies \begin{equation} \label{f_assumptions} f(1) = 1 \qquad \text{and} \qquad \frac{{\rm d}f(x)}{{\rm d}x}\Bigg|_{x=1} = -1. \end{equation} If the particle diameter $d$ is much smaller than the typical distance $\ell$ between the particles, the function $f(d/\ell)$ may be expanded to linear order, \begin{equation} \label{Taylor} f(x) = c_0 + c_1 x + \ldots \end{equation} In linear approximation the coefficients are determined by (\ref{f_assumptions}): $c_0 = 2$ and $c_1 = -1$. This determines the energy barrier (\ref{Phi}). In 1969 Salpeter and Van Horn~\cite{Salpeter69,Slattery80} pointed out, that inside a strongly coupled OCP a short-range body centered cubic (BCC) ordering will emerge. In the BCC lattice the nearest neighbour distance $\ell$ is related to the volume fraction $\nu$ by \begin{equation} \frac{d}{\ell} = \frac{2}{\sqrt{3}} \left(\frac{3}{\pi} \nu \right)^{1/3} \approx 1.14\,\nu^{1/3}. \end{equation} Assuming a BCC structure and using the linear approximation for $f(x)$ in (\ref{Phi}), the effective energy barrier is therefore given by \begin{equation} \label{E_eff_bcc} E_{\rm eff} = E_{\rm q} \left(1-2.27\,\nu^{1/3} + 1.29\,\nu^{2/3}\right) \end{equation} \begin{figure}[tb] \centerline{\psfig{figure=FIG3.ps,angle=270,width=8cm}} \caption{The dependency of the effective energy barrier on the solid fraction. Filled circles correspond to computer simulations, the solid line is Eq.~(\ref{E_eff_bcc}).} \label{work.fig.3} \end{figure} To test Eq.~(\ref{E_eff_bcc}) we simulated systems with densities ranging from $\nu = 0.001$ to $\nu = 0.216$ and determined the ratio $E_{\rm eff} (\nu) / E_{\rm q}$ as in Fig.~\ref{work.fig.2}. The results are plotted in Fig.~\ref{work.fig.3}. The agreement of the theoretical formula (\ref{E_eff_bcc}) with the simulations is excellent. One can see, that in the dilute limit the effective energy barrier extrapolates to $E_{\rm q}$. We cannot simulate low density systems, because collisions are too unlikely. For the highest densities one cannot expect that the linear approximation (\ref{Taylor}) remains valid. Also, the dense packing of spheres is achieved with an FCC (face centered cubic) rather than a BCC ordering. This may be responsible for the systematic slight deviation from the theoretical curve in Fig.~\ref{work.fig.3}. The reduction of the Coulomb repulsion was also found in the OCP, when it was applied to dense stars~\cite{Salpeter69}. There the analogue of the second term in (\ref{Phi}) is called the ``screening potential'' (somewhat misleadingly, as there is no polarizable counter charge and hence no screening). Monte Carlo simulations~\cite{Brush} of the OCP were interpreted in terms of a linear ``screening potential''~\cite{DeWitt73}, which corresponds to (\ref{Taylor}), and the analogue of the conditions (\ref{f_assumptions}) also occurs in the plasma context~\cite{Itoh}, although based on a different physical reasoning. Corrections to the linear approximation are the subject of current research~\cite{Rosenfeld}. However, applying these more sophisticated forms of the ``screening potential'' of the OCP model to dense charged granular gases seems arguable as for higher densities the influence of the hard spheres become more and more important and so the analogy to the OCP model, which uses point charges, does no longer hold. \section{Discussion} We derived the dissipation rate of a charged granular gas in the dilute limit. Compared to the uncharged case it is exponentially suppressed by a Boltzmann factor depending on the ratio between the Coulomb barrier and the granular temperature. This result was obtained assuming a Gaussian velocity distribution, although it is known that in the uncharged case deviations from a Gaussian behaviour emerge due to the inelastic collisions~\cite{Esipov}. These deviations, however, were shown to have little effect on the dissipation rate~\cite{Noije98}. As the system becomes less dissipative in our case, it is reasonable to expect that the effect of deviations from a Gaussian velocity distribution will be even weaker. One may say that a dilute granular gas with monopolar charging is more similar to a hard sphere gas in thermal equilibrium than a neutral one. In a dense system particle correlations enter the collision statistics and hence the dissipation rate in two ways: First there is the well known Enskog correction as in the uncharged case. It describes that the excluded volume of the other particles enhances the probability that two particles are in contact. Second the Coulomb barrier which colliding particles must overcome is reduced and will vanish in the limit of a dense packing. In our simulation we did not observe the {\em shearing\/} or the {\em clustering\/} instability, probably because our systems were rather small. It is reasonable to expect, however, that shearing or clustering instabilities may at most exist as transients in the presence of monopolar charging, because the Coulomb repulsion will homogenise the system in the long run. \section*{Acknowledgements} We thank Lothar Brendel, Alexander C. Schindler and Hendrik Meyer for useful comments. We gratefully acknowledge support by the Deutsche Forschungsgemeinschaft through grant No.~{Wo~577/1-2}.
\section{introduction} The non-equilibrium nature of the spin glass phase has been extensively investigated since it was discovered in low frequency ac-susceptibility measurements \cite{Lgrac} and dc-relaxation experiments in the early 1980's \cite{Lgr}. The interpretation of the experimental results and the design of new experimental procedures emanate essentially from two different theoretical approaches: hierarchical phase space models \cite{Saclay} and real space droplet/domain models \cite{three,henke}. A high level of phenomenlogical and theoretical insight into the phenomena has by now been acquired. There remain, however, unresolved problems as to the interpretation and reproducibility of non-equilibrium results. These shortcomings do in part also prevent a useful judgement in-between the applicability of the different models to real 3d spin glasses. The recently reported memory effect\cite{one,two}, observable in low frequency ac-susceptibility experiments, not only elucidates some paradoxal features of the spin glass phase, but an in detail study of related phenomena \cite{four} also emphasizes the importance of: cooling/heating rates, wait times, thermalisation times etc., i.e the detailed thermal history of the sample on the results from low frequency ac-susceptibility and dc-magnetic relaxation experiments. These complicated non-equilibrium phenomena should be considered in the perspective that an ac-experiment at only a couple of decades higher frequency, $f\geq$ 100 Hz (observation time 1/$\omega\leq$ 2 ms), in spite of a strong frequency dependence, simply shows an equilibrium character when measured in ordinary ac-susceptometers. The non-equilibrium processes of the system are only active on observation time scales governed by the cooling/heating rate and during a halt at constant temperature these evolve with the time the sample has been kept at constant temperature. In this paper we report results from low field zero field cooled (zfc) relaxation experiments, i.e. measurements of the response function, at one specific temperature in the spin glass phase. The parameter we vary in a controlled \begin{figure} \centerline{\hbox{\epsfig{figure=fig1.eps,width=8.5cm}}} \caption{ \hbox {Fc- and zfc-magnetisation of Ag(11 at$\%$Mn) }plotted vs. temperature in an applied field of 40 Oe. } \label{fig1} \end{figure} way is the thermal history in the spin glass phase. The results imply that the reponse function is governed by: the cooling procedure in a rather narrow region just above the measurement temperature, the wait time before the response function is measured, and the cooling/heating procedure in a rather narrow region just below the measurement temperature if an additional undercooling has been carried out. It is also clearly shown that the thermal history, in the spin glass phase, at temperatures well enough separated from the measurement temperature is irrelevant to the reponse function at $T_{m}$. The results are put in relation to a phenomenological real space fractal domain picture that is introduced and discussed in more detail elsewhere \cite{four,interference}. The investigation is also motivated by a lack of agreement in the detailed behaviour of the non-equilibrium dynamics, when measured on one and the same spin glass material in different magnetometers. \begin{figure} \centerline{\hbox{\epsfig{figure=fig2.eps,width=8.5cm}}} \caption{ \hbox {The relaxation rate, $S(t)=1/h$ $\partial m$$/$$\partial$ln$t$} vs. time, measured at $T_{m}$ = 27 K for three different wait times, $t_{w}$, as indicated in the figure, using two different cooling rates 3K/min. and 0.5 K/min, $h$=1 Oe. } \label{fig2} \end{figure} \section{experimental} The sample is a bulk piece of Ag(Mn 11at$\%$) with a spin glass temperature $T_{g}\approx$ 35 K made by drop-synthesis technique. The experiments were performed in a non-commercial SQUID magnetometer optimised for dynamic studies in low magnetic fields \cite{five}. The sample is cooled in zero magnetic field from a temperature above $T_{g}$, using a controlled thermal sequence, to the measurement temperature, $T_{m}$= 27 K, where a weak magnetic field ($h$= 1 Oe) is applied after the system been kept at $T_{m}$ a certain wait time, $t_{w}$. The relaxation of the magnetisation is then recorded as a function of time elapsed after the field application. In our figures of the relaxation data we show the relaxation rate, i.e. the logarithmic derivative of the response function $S$=1/$h$ d$m$/dlog$t$, which is the quantity that most clearly exposes changes of the response function after the different thermal procedures. The relaxation rate is related to the out-of-phase component of the ac-susceptibility via $S$($t$)$\approx$-2/$\pi$ $\chi''$($\omega$) at $t$=1/$\omega$ \cite{six}, and $\chi''(T)$ is also the quantity that most instructively has been used to visualise the memory phenomenon in spin glasses mentioned above \cite{two,four}. To acquaint with the sample, the field cooled and zero field cooled magnetisation is plotted vs. temperature in Fig. 1. The curves are measured in a field of 40 Oe. The cusp in the zfc susceptibility around 35 K closely reflects the spin glass temperature $T_{g}$ of the sample. \section{Results} The classic aging experiment is performed by cooling the sample directly to a measurement temperature below $T_{g}$, wait a controlled amount of time and then switch the magnetic field on or off. Not much attention has been given to the influence of the cooling rate. In Fig. 2, the relaxation rate, $S(t)$, is plotted vs. log$t$ for three different wait times. The results for two substantially different, but still in a logarithmic time perspective rather similar, cooling rates are plotted, 3 K/min open symbols and 0.5 K/min solid symbols. The wait time dependence of the data displays the characteristics of the ageing phenomenon in spin glasses (a similar ageing phenomenon is also an inherent property of other disordered and frustrated magnetic systems \cite{seven}). The response is visually sensitive to the cooling rate, $t_{c}$, but the influence of $t_{c}$ decreases as the wait time increases. For the $t_{w}$=10 s curves, rapid cooling yields a maximum in $S$($t$) at a shorter observation time than slow cooling. The two curves are clearly different and do not coalesce anywhere in the measured time interval. For the longer wait times, the position of the maxima do not differ appreciably, but the magnitude is somewhat higher for the rapidly cooled curves. At short observation times, for the $t_{w}$=$10^3$ and $10^4$ s curves, there are no or weak differences in the relaxation rates, but after some time they start to deviate and there remains a cooling rate dependence all through the long time part of our observation time window. Concentrating on the position of the maximum in the relaxation rate and identifying this with an effective age of the system, $t_{aeff}$, this parameter is apparently governed by the cooling rate and the wait time. When the wait time is short, $t_{aeff}$ is governed by the cooling rate, whereas for longer wait times, it is dominated by the wait time, $t_{w}$. The position of the maximum in the relaxation rate is closely equal to the wait time for $t_{w}$ = $10^3$ and $10^4$ s while for the $t_{w}$ = 10 s the maximum is delayed about a decade in time. Similar tendencies as to the evolution of the effective age with increasing wait time have earlier been reported in connection with a specific method, `field quenching', to achieve a well defined initial state for ageing experiments \cite{fieldquenching}. To further investigate how the thermal sequence on approaching the measurement temperature influences the response function, temperature shift experiments under controlled cooling and heating rates were performed. In the negative temperature shift experiment, the system is kept a certain wait time at a shift temperature, $T_{s}$, above $T_{m}$. Therafter the system is cooled to the measurement temperature where the field is applied immediatly after the sample has become thermally stabilised. The positive temperature shift experiment follows a similar procedure, but the temperature $T_{s}$ is below $T_{m}$. Fig. 3 shows the relaxation rate when the system has been subjected to (a) negative and (b) positive temperature shifts. The wait time at $T_{s}$ is 1000 s and $T_{m}$ = 27 K. The cooling and heating rates are 0.5 K/min. Two reference curves measured after the sample has been cooled at 0.5 K/min directly to $T_{m}$ are included in the figure. One corresponds to a wait time $t_{w}$ = 1000 s and the other is measured without any wait time prior to the field application ($t_{w}\approx$ 0). From Fig. 3(a) it is seen that for $T_{s}$ $>$ 30 K the response is indistinguishable from what is measured if the sample \begin{figure} \centerline{\hbox{\epsfig{figure=fig3.eps,width=8.5cm}}} \caption{ \hbox {The relaxation rate, $S(t)$ vs. time, employing a} halt lasting 1000 s in the cooling procedure at different temperatures (a) $T_{s}$ = 34, 32, 30, 29, 28 and 27.5 K above $T_{m}$ = 27 K and at different temperatures (b) $T_{s}$ = 15, 20, 24, 25, 26 and 26.5 K below $T_{m}$ = 27 K (b). In both (a) and (b) two reference curves are included that are measured after immediately cooling the sample to $T_{m}$ at a cooling rate of 0.5 K/min. and using $t_{w}\approx$ 0 and 1000 s, $h$= 1 Oe. } \label{fig3} \end{figure} is cooled directly to $T_{m}$. The time the sample has been kept at $T_{s}$ is irrelevant for the response at $T_{m}$. When the shift temperature $T_{s}$ is lower than 30 K, the time spent at the higher temperature starts to influence the response at $T_{m}$. The first deviation from the reference ($t_{w}\approx$ 0) curve occurs at long observation times, the maximum in the relaxation rate decreases in magnitude and on further decreasing $T_{s}$ it shifts towards longer times, and finally when $T_{s}$ approaches $T_{m}$, the relaxation rate approaches the $t_{w}$ = 1000 s reference curve. The important implication of this behaviour is that only the thermal history in a limited temperature region just above $T_{m}$ governs the reponse function. The results from positive temperature shifts shown in Fig. 3(b) give a somewhat more complicated result. A first observation is that when the system has been cooled to and aged at temperatures well below $T_{m}$, the measured curves are identical, but different from the $t_{w}\approx$ 0 reference curve, the maximum in the relaxation rate \begin{figure} \centerline{\hbox{\epsfig{figure=fig4.eps,width=8.5cm}}} \caption{ \hbox {The relaxation rate $S(t)$ vs. time, after cooling} the sample to a low temperature $T_{s}$ = 20 and 24.5 K, halt at this temperature for 0 s or 1000 s and then approach $T_{m}$ = 27 K at a heating rate of 0.5 K/min. where the field is immediately applied. A reference curve $t_{w}\approx$ 0 measured after cooling the sample directly to $T_{m}$ without any halt is also plotted, $h$ = 1 Oe } \label{fig4} \end{figure} occurs at a shorter time than for the reference curve. Thus, the response function is somewhat different after only cooling to the measurement temperature at one specific cooling rate, compared to after substantially undercooling the sample and re-heat it with a similar heating rate. The curves representing $T_{s}$= 15 and 20 K are indistiguishable from each other, but when $T_{s}$ is increased further, the relaxation rate gets surpressed and the maximum shifts towards longer times to finally coalesce with the reference $t_{w}$= 1000 s curve when $T_{s}\approx$ $T_{m}$. Fig. 4 shows relaxation rate curves from experiments where the sample is undercooled to $T_{s}$, and either kept there 1000 s (as in Fig. 3(b)) or immediately re-heated to $T_{m}$, where the relaxation is measured using $t_{w}\approx$ 0. The cooling/heating rates are again 0.5 K/min. The figure shows that the wait time at the lowest temperature does not affect the results, identical curves are measured whether the sample is kept at 20 K for 1000 s or 0 s. For a shift temperature, $T_{s}$= 24.5 K, closer to $T_{m}$, a clear difference is observed between the two curves with different wait times. The implication from these data is again that the thermal history far enough away from the measurement temperature is irrelvant to the response function at $T_{m}$, the response is in the experimental time window fully governed by the previous cooling/heating history in the very neighbourhood of $T_{m}$. In Fig. 5, results for (a) negative and (b) positive temperature shifts using two different cooling rates 3 K/min and 0.5 K/min (as in Fig. 3) are displayed. The cooling/heating rate from $T_{s}$ to $T_{m}$ = 27 K is always the same, 0.5 K/min. For negative temperature shifts (Fig. 5 (a)) there is no influence of the different cooling rates for $T_{s}$ $>$ 29 K. The rapid and slow cooling give the same response at the measurement temperature. For $T_{s}$ = 28 K influences of the different cooling rates start to become observable at observation times, $t$ $>10^3$ s. At shorter observation times no sign of the different cooling rates can be resolved. When $T_{s}$ gets even closer to $T_{m}$, the influence from the differences in cooling rates appears earlier, but still there is a region at shorter observation times where the response is independent of the cooling rate. For positive temperature shifts, Fig. 5 (b), the result is different. There are clear cooling rate effects for all different temperatures $T_{s}$. The curve representing rapid cooling always has a larger magnitude and a sharper maximum that occurs at shorter observation times than the corresponding slowly cooled curve. This result implies that the cooling rate in the region just above $T_{m}$ always remains one of the governing parameters for the response after undercooling the sample, i.e. a memory of this cooling process becomes imprinted in the spin structure and is conserved in spite of the re-structuring that occurs at lower temperatures. \section{Discussion} Our current results on the non-equilibrium response function show that the cooling rate significantly influences the measured response function, especially the reponse at short wait times is dominated by the specific cooling rate employed. Also, in experiments where the sample has been undercooled below the measurement temperature, the cooling rate above the measurement temperature remains one of the governing parameters for the non-equilibrium response. These findings have practical importance for the design of experimental procedures and when making detailed comparisons between results obtained in different experimental set-ups and in different laboratories on similar spin glass materials. We have, using different experimental procedures, elucidated an apparently paradoxal property of the non-equilibrium spin structure in spin glasses: the spin structure records the cooling history, this thermal history (cooling/heating rate, wait times etc.) remains imprinted in the configuration, and the fragment of the thermal history confined in a rather narrow region close to $\it any$ higher temperature $T_{m}$ within the spin glass phase, can be recovered in a relaxation experiment at $T_{m}$. A continuous memory recording occurs on cooling, in spite of the fact that the spin structure is subjected to substantial reconfiguration at all temperatures below $T_{g}$ as is observed from the ever present ageing behaviour; and although the short time response appears equilibrated at all temperatures in the spin glass phase. Employing the droplet scaling model \cite{three} and the concepts chaos and overlaplength \cite{eight}, the observations of non-equilibrium and ageing behaviour observed after cooling the sample to a temperature in the spin glass phase can be accounted for as an immediate consequence of the growth of the size of correlated spin glass domains at \begin{figure} \centerline{\hbox{\epsfig{figure=fig5.eps,width=8.5cm}}} \caption{ \hbox {The relaxation rate, $S(t)$ vs. time, cooling the} sample at two different rates, 3 K/min. and 0.5 K/min., employing a halt lasting 1000 s in the cooling procedure at different high temperatures (a) $T_{s}$ = 34, 29, 28 and 27.5 K before approaching $T_{m}$ = 27 K with a cooling rate of 0.5 K/min., and at different low temperatures (b) $T_{s}$ = 15, 22, 24, 25 and 26 K before approaching $T_{m}$ = 27 K with a heating rate of 0.5 K/min. The magnetic field $h$=1 Oe is applied immediately when reaching $T_{m}$. } \label{fig5} \end{figure} constant temperature. Interpretations of ageing in these terms have been extensively discussed in numerous articles and are recently reviewed in ref.\cite{one}. A somewhat modified version of the droplet scaling model that has been constructed to account for the memory behaviour\cite{two} is extensively discussed in ref. \cite{four}. The continuous memory recording of the thermal history that is exemplified by the current experimental results does however require a comment on the paradoxal possibility that reconfiguration on different length scales are fully separable, i.e. that reconfiguration of the spin structure on short length scales allows a simultaneous stability of the old configuration on larger length scales. The droplet model prescribes one unique equilibrium spin glass state with time reversal symmetry and that domains of spin glass ordered regions grow unrestrictedly with time at constant temperature. If the temperature is altered, the equilibrium configuration also alters due to chaos, but there is an overlap on short length scales between the equilibrium configurations, the length scale of which rapidly decreases with increasing separation between the two temperatures. The experimental results show that a memory of the spin structure that has developed at high temperatures remains imprinted in the non-equilibrium structure at lower temperatures (but also that it is rapidly erased if the temperature is increased above the original aging temperature). Such a re-stored spin structure requires that all reconfiguration at lower temperatures must occur only on small lengths scales and in dispersed regions. The bulk of the numerous droplet excitations of different sizes that do occur may not cause irreversible changes of the spin structure on large length scales, but are to be excited within already equilibrated regions of spin glass order. How can this rather abstract picture of the dynamic spin structure be related to our current experimental observations? The processes that cause the increase of the magnetisation in a zfc magnetic relaxation experiment is a polarisation of spontaneous droplet excitations. The measured quantity, the zfc magnetisation, gives an integrated value of polarisation of all droplet excitation with relaxation time shorter than the observation time and the measured relaxation corresponds to droplets with relaxation time of the order of the observation time, $t$. In an ac-susceptibility experiment, the in-phase component gives the integrated value of all droplet excitations with relaxation time shorter than the observation time, $t$ = 1/$\omega$, and the out-of phase component measures the actual number of droplets of relaxation time equal to 1/$\omega$. The non-equilibrium characteristics imply that the distribution of droplet excitations changes with the time spent at constant temperature and that there is an excess of droplet excitations of a size that corresponds to a relaxation time of the order of the wait time. On shorter time scales an equilibrium distribution has been attained, reflected by the equilibrium reponse always obtained in ac-susceptibility experiments at higher frequencies, and in zfc measurements at long but different wait times by the fact that a similar (equilibrium) reponse is approached at the shortest observation times. The implication of the cooling rate dependence is that the actual distribution of active droplets is governed by the cooling rate and the wait time at constant temperature, and that this distribution in turn is governed by the underlying spin configuration. The phenomenon that the sample retains a distribution of droplets that is governed by the cooling rate and any previous wait time at the measurement temperature, then implies that a closely equivalent spin configuration to the original one is also retained when the temperature is recovered. \section{conclusions} We have shown that non-equilibrium dynamics measured at a specific temperature in spin glasses is primarily governed by the thermal history close to this temperature during the cooling sequence. If the sample has been undercooled, the response is also partly affected by the heating rate towards the measurement temperature. The behaviour may be incorporated in a real space picture of a random spin configuration containing fractal spin glass domains of sizes that increase through spontaneous droplet excitations. The results emphasise the importance of well controlled experimental procedures when studying non-equilibrium dynamics of spin glasses. \section{acknowledgments} Financial support from the Swedish Natural Science Research Council (NFR) is acknowledged. Numerous and useful discussions on the memory phenomenon and the non-equilibrium nature of the spin glass phase with T. Jonsson, E. Vincent, J.-P. Bouchaud and J. Hammann are acknowledged. \begin {references} \bibitem{Lgrac} L. Lundgren, P. Svedlindh and O. Beckman, J. Magn. Magn. Mater. {\bf 31-34}. 1349 (1983). \bibitem{Lgr} L. Lundgren, P. Svedlindh, P. Nordblad and O. Beckman, Phys. Rev. Lett. {\bf 51}, 911 (1983). \bibitem{Saclay} see e.g. E. Vincent, J.P. Bouchaud, J. Hammann and F. Lefloch, Phil. Mag. B {\bf 71} (1995); J.-P. Bouchaud, L.F. Cugliandolo, J. Kurchan and M. M\'ezard, in {\it Spin Glasses and Random Fields} p 161-223, ed. A. P. Young, World Scientific (1998). \bibitem{three} D. S. Fisher and D. A. Huse, Phys. Rev. B {\bf 38}, 373 (1988); {\bf 38}, 386 (1988). \bibitem{henke} G.J.M. Koper and H.J. Hilhorst, J. Phys. (France) {\bf 49}, 429 (1988). \bibitem{one} P. Nordblad and P. Svedlindh; in {\it Spin Glasses and Random Fields} p 1-27, ed. A. P. Young, World Scientific (1998). \bibitem{two} K. Jonason, E. Vincent, J. Hammann, J. P. Bouchaud, P. Nordblad, Phys. Rev. Lett. {\bf 81}, 3243 (1998). \bibitem{four} T. Jonsson, K. Jonason, P. J\"{o}nsson and P. Nordblad, Phys. Rev. B {\bf 55}, 8770 (1999). \bibitem{interference} K. Jonason, P. Nordblad, E. Vincent, J. Hammann and J.P. Bouchaud, unpublished. \bibitem{five} J. Magnusson, C. Djurberg, P. Granberg, P. Nordblad Rev. Sci. Instrum. {\bf 68}, 3761 (1997). \bibitem{six} L. Lundgren, P. Svedlindh and O. Beckman, J. Magn. Magn. Mater. {\bf 25}, 33 (1981). \bibitem{seven} P. Nordblad; in {\it Dynamical Properties of Unconventional Magnetic Systems} p 343-366, eds. A.T. Skjeltorp and D. Sherrington, Kluwer (1998). \bibitem{fieldquenching} P. Nordblad, P. Svedlindh, P. Granberg and L. Lundgren, Phys. Rev. B {\bf 35}, 7150 (1987). \bibitem{eight} A. J. Bray and M. A. Moore, Phys. Rev. Lett. {\bf 58}, 57 (1987). \end{references} \end{multicols} \end{document}
\section{Introduction} \noindent Random matrix theory is actively developing. Among numerous topics of the theory and its various applications those related to the asymptotic eigenvalue distribution of random matrices of large order are of considerable interest. An important role in this branch of the theory plays the eigenvalue counting measure defined for any Hermitian or real symmetric matrix \begin{equation} M_{n}=\left\{ M_{jk}^{(n)}\right\} _{j,k=1}^{n} \label{mat} \end{equation} as follows \begin{equation} N_{n}(\Delta )= \frac{1}{n} \natural \left\{ \lambda _{i}^{(n)}\in \Delta \right\} \label{N} \end{equation} where $\Delta $ is a Borel set of the real axis $\mathbf{R}$ and $\{\lambda _{i}^{(n)}\}_{i=1}^{n}$ are eigenvalues of $M_{n}$. One distinguishes several large-$n$ asymptotic regimes for the probability properties of eigenvalues (see e.g. \cite{Me:91}). In this paper we deal with the global regime, defined by the requirement that \ the expectation $\mathbf{E}(N_{n}(\Delta ))$ has well defined (i.e. not zero and not infinite) weak limit \begin{equation} N(\Delta )=\lim_{n\rightarrow \infty }\mathbf{E}(N_{n}(\Delta )). \label{IDS} \end{equation} This limit is called the Integrated Density of States (IDS). We shall see below that explicit conditions to be in the global regime may look differently in different cases. The IDS is a quantity to be found and analyzed first in any random matrix study, because it enters in practically any problem and result of the theory. A new wave of interest to the global regime is motivated by recent studies of the free group factors of operator algebras known now as free probability theory \cite{Vo-Di-Ni:92}. In this paper we present a simple approach of the study of the random measure (\ref{N}) in the global regime. The approach allows us to find the limit (\ref{IDS}) in many interesting cases, to show that the sequence of random measures (\ref{N}) converges to this nonrandom limit either in probability or even with probability 1 and to find bounds on the rate of the convergence. The paper is organized as follows. In Section 2 we consider most studied random matrix ensembles, in particular, the Gaussian and the Circular Ensembles. This allows us to explain the method by using the simple and known setting of these ensembles. We call these ensembles classical because, first, they are indeed classic objects of the theory, and second, because of the role of the classical orthogonal polynomials in their studies (although, we almost do not use this technique in the paper). Section 3 is devoted to ensembles whose probability distribution is invariant with respect to unitary or orthogonal transformations and whose study is motivated by the Quantum Field Theory. In Section 4 we present new results on the form of the limiting normalized eigenvalue counting measure of the sum of two Hermitian or real symmetric matrices randomly rotated one with respect to another. In Section 5 we first discuss ensembles with independent but not necessary Gaussian entries. These ensembles are known as the Wigner Ensembles. Then we consider the ensembles that can be represented as the sum of the rank one independent operators. This form generalizes the sample covariance matrices widely used in multivariate analysis and is also motivated by statistical mechanics. Most of results, presented in the paper are known, sometimes for decades. However they were obtained by different and often rather complicated methods while in this paper we derive them in the framework of an unique approach, that we present in three slightly different versions, according to the case considered. We do not give here complete proofs of all presented results, but only outline basic moments of their proofs. The complete versions of the proofs will be published in \cite{Kh-Pa-St:99}. \section{ Classical Ensembles} \subsection{Gaussian Ensembles} We start from the well known ensembles among which the Gaussian Ensembles (GE) are most known. We restrict ourself by the technically simplest of Gaussian Ensembles, consisting of Hermitian matrices and known as the Gaussian Unitary Ensemble (GUE) because its probability distribution \begin{equation} P_{n}(dM)=Z_{n}^{-1}\exp \left(-\frac{n}{4w^{2}}\hbox{ Tr } M^{2}\right)dM \label{gue} \end{equation} is unitary invariant. Here \begin{equation} dM=\prod\limits_{1\leq j\leq n}dM_{jj}\prod_{1\leq j<k\leq n}d\mbox{Re}% M_{jk}d\mbox{Im}M_{jk} \label{mes} \end{equation} is the ``Lebesgue'' measure on the space of $n\times n$ Hermitian matrices. It is often convenient to write \begin{equation} M_{n}=n^{-1/2}W_{n} \label{w} \end{equation} where now $W_{n}=\left\{ W_{jk}\right\} _{j,k=1}^{n}$ can be considered as the left upper corner of the semi-infinite Hermitian matrix $W=\left\{ W_{jk}\right\} _{j,k=1}^{\infty }$, whose entries are complex Gaussian random variables defined by the relations \begin{equation} \mathbf{E}(W_{jk})=0, \quad \mathbf{E}(W_{jk} \overline {W_{lm}})= 2w^{2}\delta _{jl}\delta _{km}. \label{mom} \end{equation} Thus the probability space in this case consists of these matrices and has as a probability measure the infinite product of the Gaussian measures defined by relations (\ref{mom}). The relations (\ref{w}) and (\ref{mom}) define the global regime in this case. \begin{theorem} For the Gaussian Unitary Ensemble defined above the sequence of eigenvalue counting measures (\ref{N}) converges with probability 1 to the nonrandom measure \begin{equation} N_{sc}(\Delta )=\frac{1}{4\pi w ^{2}}\int_{\Delta \cap \lbrack -2 \sqrt{2} w, 2\sqrt{2}w]}\sqrt{% 8w^{2}-\lambda ^{2}}d\lambda \label{sc} \end{equation} i.e. $N_{sc}$ has the density \begin{equation} \rho _{sc}(\lambda)=(4\pi w^{2})^{-1}\sqrt{8w^{2}-\lambda ^{2}} \label{rosc} \end{equation} concentrated on the interval $[-2\sqrt{2}w,2\sqrt{2}w]$. The convergence $N_{n}$ to $N_{sc}$ has to be understood as the weak convergence of measures. \end{theorem} The theorem dates back in fact to Wigner (see \cite{Me:91}). We give below the two proofs of the theorem to illustrate two methods that can be used in rather general situation of dependent and not necessary Gaussian entries. Both proofs as well as other proofs in this paper are based on the study the \textit{Stieltjes transforms} of measures instead of measures themselves. In the random matrix theory the Stieltjes transform was used for the first time in paper \cite{Ma-Pa:67} and since then is proved to be a rather efficient tool of the study of the global regime. \bigskip Recall that the Stieltjes transform $f(z)$ of a non-negative measure $m(d\lambda ),$ $m(\mathbf{R})=1,$ is the function of the complex variable $z$ defined for all non-real $z$ by the integral \begin{equation} f(z)=\int \frac{m(d\lambda )}{\lambda -z},\text{ \ }\mbox{Im}z\neq 0. \label{St} \end{equation} Here and below we use integrals without indicated limits denote integrals over whole real axis. $f(z)$ \ is obviously analytic for non-real $z$ and satisfies the conditions \begin{equation} \mbox{Im}f\cdot \mbox{Im}z>0,\mbox{Im}z\neq 0, \quad \sup_{y\geq 1}y|f(iy)|=1. \label{Nev} \end{equation} It can be shown \cite{Ak-Gl:61} that any function $f(z)$ \ defined and analytic for non-real $z$ and satisfying conditions (\ref{Nev}) is the Stieltjes transform of a unique nonnegative and normalized to 1 measure $% m(d\lambda )$ and that for any continuous function $\varphi (\lambda )$ with a compact support \begin{equation} \int \varphi (\lambda )m(d\lambda )=\lim_{\varepsilon \rightarrow 0} \frac{1}{\pi}\int \varphi (\lambda )\mbox{Im}f(\lambda +i\varepsilon )d\lambda \label{fp} \end{equation} Besides the one-to-one correspondence between measures and their Stieltjes transforms is continuous if one will consider the weak convergence of measures and the convergence of the Stieltjes transforms that is uniform on all compacts in $\mathbf{C\backslash R}$. \bigskip The use of the Stieltjes transform in this context is based on the spectral theorem expressing the Stieltjes transform \begin{equation} g_{n}(z)=\int \frac{N_{n}(d\lambda )}{\lambda -z} \label{gn} \end{equation} of the normalized eigenvalue counting measure (\ref{N}) of a matrix $M_n$ via its resolvent \begin{equation} G(z)=(M_n-z)^{-1},\text{ \ }\mbox{Im}z\neq 0 \label{fG} \end{equation} by the formula \begin{equation} g_{n}(z)=\frac{1}{n}\hbox{ Tr } G(z) \label{gG} \end{equation} Our proof has as a basic ingredient the following \begin{proposition} The Stieltjes transform $g_{n}(z)$ of the eigenvalue counting measure (\ref{N}) for the Gaussian Unitary Ensemble defined by (\ref{gue}) or by (\ref{w}) and (\ref {mom}) has the asymptotic properties for $\mbox{Im}z\geq y_{0}=5w$: \begin{equation} \lim_{n\rightarrow \infty }\mathbf{E}(g_{n}(z))=f_{sc}(z), \label{exp} \end{equation} \begin{equation} \mathbf{E}(|\gamma _{n}(z)|^{2})\leq \frac{C}{w^{2}n^{2}} \label{var} \end{equation} where \begin{equation} \gamma _{n}(z)=g_{n}(z)-\mathbf{E}(g_{n}(z)), \label{gam} \end{equation} $f_{sc}\left( z\right) $ is the unique solution of the equation \begin{equation} 2w^{2}f^{2}+zf+1=0 \label{qua} \end{equation} verifying condition (\ref{Nev}), and we denote here and below by $C$ numerical constants that may be different in different formulas. \end{proposition} To prove the proposition we need the following elementary facts: \begin{enumerate} \item[(i)] for any two matrices $A$ and $\ B$ \begin{equation} (B-z)^{-1}=(A-z)^{-1}-(B-z)^{-1}(B-A)(A-z)^{-1}, \label{res} \end{equation} (the resolvent identity); \item[(ii)] if $\zeta $ is a complex valued Gaussian random variable defined by \begin{equation*} \mathbf{E}(\zeta )=\mathbf{E}(\zeta ^{2})=0,\text{ \ }\mathbf{E}% (|\zeta |^{2})=2w^{2} \end{equation*} and $\varphi (z,\bar{z})$ is a differentiable function polynomially growing at infinity and having the same property of its derivatives, then \begin{equation} \mathbf{E}(\zeta \varphi (\zeta,\bar{\zeta} ))=\mathbf{E}(|\zeta |^{2})% \mathbf{E}\left(\frac{\partial \varphi }{\partial \bar{\zeta}}\right)= 2w^{2}\mathbf{E}\left(% \frac{\partial \varphi }{\partial \bar{\zeta}}\right); \label{difg} \end{equation} \item[(iii)] for the resolvent $G(z)=(A-z)^{-1}$ of any Hermitian or real symmetric matrix $A$ we have \begin{equation} ||G(z)||\leq |\mbox{Im}z|^{-1},\text{ \ }|G_{jk}(z)|\leq |\mbox{Im}z|^{-1} \label{norm} \end{equation} where $G_{jk}(z),j,k=1...n$ are the matrix elements of the resolvent. \end{enumerate} Now using (\ref{res}) for the pair $B=M,A=0$ and applying (\ref{difg}) we obtain the system of identities for the moments \begin{equation} m_{p}(z_{1},...,z_{p})=\mathbf{E(}g_{n}(z_{1})...g_{n}(z_{p})\mathbf{)} \label{momp} \end{equation} of the random function $g_{n}$ with non-real arguments $z_{1},...,z_{p}$% \begin{equation} m_{1}(z_{1}) =-\frac{1}{z_{1}}-\frac{2w^{2}}{z_{1}}m_{2}(z_{1,}z_{1}), \label{sys} \end{equation} \begin{equation*} m_{p}(z_{1},...,z_{p}) =-\frac{1}{z_{1}}m_{p-1}(z_{2},...,z_{p})-\frac{% 2w^{2}}{z_{1}}m_{p+1}(z_{1},z_{1,}z_{2,}...,z_{p})+r_{p}(z_{1},...,z_{p}),p% \geq 2. \end{equation*} where \begin{equation*} r_{p}(z_{1},...,z_{p})=\frac{2w^{2}}{n^{2}z_{1}}\sum_{q=2}^{p}\mathbf{E(}% \frac{1}{n}% \hbox{ Tr } (G_{n}(z_{1})G_{n}^2(z_{q}))g_{n}(z_{2})...g_{n}(z_{q-1}) g_{n}(z_{q+1})...g_{n}(z_{p}), \mathbf{)} \end{equation*} Assume now that $|\mbox{Im}z_{q}|\geq y,q=1,...$ for some $y>0$. Then relations (\ref{gG}) and (\ref{norm}) imply the bounds \begin{equation} |g_{n}(z)|\leq \frac{1}{|\mbox{Im}z|}\leq \frac{1}{y}, \label{gest} \end{equation} \begin{equation*} |m_{p}|\leq \frac{1}{y^{p}},\quad |r_{p}|\leq \frac{2w^{2}p}{n^{2}y^{p+1}}. \end{equation*} Following statistical mechanics (see e.g.\cite{Ru:69}) we can treat system of identities (\ref{sys}) as a linear equation in the Banach space $\mathcal{% B}$ of complex valued sequences $m$ of functions $m=\{m_{p}(z_{1},...,z_{p})\}_{p=1}^{\infty }$ equipped with the norm \begin{equation} ||m||=\sup_{p\geq 1}\eta ^{p}\sup_{|\mbox{Im}z_{q}|\geq y,q=1,..p} |m_{p}(z_{1},...,z_{p})| \label{Bnorm} \end{equation} for some $\eta >0$. The equation has the form $\ $% \begin{equation} m=Am+b+r\bigskip \ \label{Eq} \end{equation} where \begin{eqnarray*} (Am)_{1}(z_{1}) &=&-\frac{2w^{2}}{z_{1}}m_{2}, \\ (Am)_{p}(z_{1},...,z_{p}) &=&-\frac{2w^{2}}{z_{1}}m_{p-1}(z_{2},...,z_{p-1})-\frac{2w^{2}}{% z_{1}}m_{p+1}(z_{1},z_{1},z_{2},...,z_{p-1}),p\geq 2, \end{eqnarray*} and $b=(-z_{1}^{-1},0,...), \quad \ r=\{r_{p}\}_{p=1}^{\infty }$. It is easy to show that optimal with respect to $\eta$ bound for the norm of $A$ is $% ||A||\leq 2^{\frac{3}{2}}$ $wy^{-1}$ for $\eta =\sqrt{2}w$ . Choosing say $% y\geq 5w$ we have uniformly in $n$ that the norm of $A$ is strictly less than 1, the vectors $m,b$ and $r$ belong to $\mathcal{B}$ and \begin{equation} ||r||\leq \frac{C}{n^{2}}. \label{rem} \end{equation} Thus equation (\ref{Eq}) is uniquely soluble uniformly in $n$. It is easy to check that this equation with $r$ replaced by zero has the factorized solution \begin{equation*} m_{p}^{(0)}(z_{1},...,z_{p})=\prod_{q=1}^{p}f(z_{q}) \end{equation*} where $f$ verifies (\ref{qua}). Thus,\bigskip\ in view of (\ref{rem}) we have \begin{equation} |m_{p}(z_{1},...,z_{p})-\prod_{q=1}^{p}f(z_{q})|\leq \frac{C}{w^{p}n^{2}} \label{fact} \end{equation} uniformly in \begin{equation} |\mbox{Im}z|\geq 5w. \label{semip} \end{equation} In particular we obtain that \begin{eqnarray*} \lim_{n\rightarrow \infty }m_{1}(z) &=&f(z), \\ |m_{2}(z,\overline{z})-m_{1}(z)m_{1}(\overline{z})| &\leq &\frac{C}{% n^{2}w^{2}} \end{eqnarray*} Recalling definition (\ref{momp}) of $m_{p}$ we see that the first relation implies that $f(z)$ satisfies (\ref{Nev}). The unique solubility of (\ref {qua}) in this class can be easily verified. The second relation is just another form of (\ref{var}). The proposition is proved. \bigskip \textbf{Remark}. As was mentioned the technique of the proof is similar to the technique of the correlation equations of the statistical mechanics (the Kirkwood-Salzburg equations, the Montroll-Mayer equations, etc. \cite{Ru:69}) combined with the mean field approximation also widely used in the statistical mechanics.The reason to have here an analogue of the mean field approximation regime is again similar to that of statistical mechanics: the entries of the GUE matrices are all of the same order of magnitude (see (\ref{w}) - (\ref{mom})), like interactions in the Curie-Weiss model. Thus to obtain a nontrivial (not zero and not infinite) limit (\ref{IDS}) (an analogue of an extensive quantity per unit volume in the statistical mechanics) we have to introduce the $n$% -dependent normalizing factor $n^{-1/2}$ in (\ref{w}). This fixes the global regime scaling but also leads to vanishing of the statistical correlations and to the factorization of the moments (see (\ref{var}) and (\ref{fact})) and to a nonlinear self-consistent equation determining the first moment that can be regarded as an analogue of the Curie-Weiss equation for the magnetization. \bigskip To prove Theorem 2.1 denote by $N_{sc}$ the measure (\ref{N}) corresponding to $f_{sc}(z)$ via (\ref{fp}). In view of Proposition 2.1 and the Borel-Cantelli lemma we have for any fixed $z$ with $|\mbox{Im}z|\geq 5w$ the convergence of $g_{n}(z)$ to $f_{sc}(z)$ with probability 1 for any fixed $z$. Since any analytic function is uniquely determined by its values on a countable set having at least one accumulation point we have the convergence of $g_n$ to $f$ with probability 1 on any compact of the domain $|\mbox{Im}z|\geq 5w$. Now by continuity of the correspondence between measures and their Stieltjes transforms we obtain that the measure $N_{n}$ converges weakly to $N_{sc}$ with probability 1. Solving explicitly equation (\ref{qua}) in the class (% \ref{Nev}) we find that \begin{equation} f_{sc}(z)=\frac{1}{4w^{2}}(\sqrt{z^{2}-8w^{2}}-z) \label{fsc} \end{equation} where the radical is defined by the condition that it behaves as $z$ as $% z\rightarrow \infty $. By using the inversion formula (\ref{fp}) we obtain\thinspace\ (\ref{sc}). Theorem 2.1 is proved. \bigskip The method was proposed in \cite{Pa-Kh:89} and subsequently used in \cite {Kh-Mo-Pa:92,Kh-Kh-Pa-Sh:92,Kh-Pa:93,Kh-Ki-Pa:94} to study a variety of problem of random matrix theory and its applications. A certain disadvantage of the method is that it is rather tedious. An important simplification of the method was proposed by A.Khorunzhy \cite{Kh:96}. We describe now this simpler version by giving another proof of the previous proposition.\bigskip\ Rewrite the first two equations of the system (\ref{sys}) \ for $% z_{1}=z,z_{2}=\bar{z}$ in the form \begin{equation} \mathbf{E}(g_{n}(z))=-\frac{1}{z}-\frac{2w^{2}}{z}\mathbf{E}(g_{n}^{2}(z)), \label{m1} \end{equation} \begin{equation} \mathbf{E}(|g_{n}(z)|^{2})=-\frac{1}{z}\mathbf{E}(\overline{g_{n}(z)})-\frac{% 2w^{2}}{z}\mathbf{E}(g_{n}^{2}(z)\overline{g_{n}(z)})+r_{2}(z). \label{m2} \end{equation} Expressing $-1/z$ in the first term of the r.h.s. of (\ref{m2}) from (\ref{m1}) we obtain \begin{equation} \mathbf{E}(|\gamma _{n}(z)|^{2})=-\frac{2w^{2}}{z}\mathbf{E}% (g_{n}^{2}(z)\overline{\gamma _{n}(z)})+r_{2}(z) \label{varga} \end{equation} where $\gamma _{n}(z)$ is defined in (\ref{gam}). By using this definition we can rewrite the expectation in the first term in the r.h.s. of this relation as $\mathbf{E}(g_{n}(z)|\gamma _{n}(z)|^{2})+\mathbf{E}(g_{n}(z))% \mathbf{E}(|\gamma _{n}(z)|^{2})$. Hence in view of (\ref{norm}) \begin{equation} |\mathbf{E}(g_{n}^{2}(z)\overline{\gamma _{n}(z)})|\leq \frac{2}{|\mbox{Im}z|}% \mathbf{E}(|\gamma _{n}(z)|^{2}) \label{ggamma}. \end{equation} By using this inequality and (\ref{gest})\ we get from (\ref{varga}) \begin{equation} (1-\frac{4w^{2}}{|\mbox{Im}z|^{2}})\mathbf{E}(|\gamma _{n}(z)|^{2})\leq \frac{2w^{2}}{|\mbox{Im}z|^{4}n^{2}}. \label{prevar} \end{equation} Thus, we obtain the bound (\ref{var}) on the variance of $g_{n}(z)$ under condition (\ref{semip}). By using this bound we replace $\mathbf{E}(g_{n}^{2}(z))$ in (\ref{m1}) by $% \mathbf{E}^{2}(g_{n}(z))$ with an error of the order $O(1/n^{2})$ uniformly in the domain (\ref{semip}). Now by using standard compactness arguments we can prove that any subsequence of the sequence $\mathbf{E}(g_{n}(z))$ converges uniformly on compacts of the domain (\ref{semip}) to a solution of equation (\ref{qua}) satisfying (\ref{Nev}). Since this solution is unique we obtain other assertions of the Proposition 2.1. \bigskip \textbf{Remarks}. (i). Analogous results are valid for two other widely used ensembles, the Gaussian Orthogonal Ensemble (GOE) consisting of real symmetric matrices distributed according the real analogue of (\ref{gue}), and for the Gaussian Symplectic Ensemble (GSE), consisting of self-dual Hermitian matrices having also the Gaussian probability distribution (see \cite{Me:91} for definitions and properties). This allows us to write that the density of the semicircle law of all three cases is concentrated on the interval $[-2\sqrt{\beta }w,2\sqrt{\beta }w]$ and has the form \begin{equation} \rho _{sc\beta }(\lambda )=\frac{1}{2\beta \pi w^{2}}\sqrt{4\beta w^{2}-\lambda ^{2},} \quad |\lambda |\leq 2\sqrt{\beta }w. \label{robeta} \end{equation} for the GOE ($\beta =1$), the GUE ($\beta =2$), and the GSE ($\beta =4$) cases respectively. (ii) We can consider an ensemble of the more general form \begin{equation} H_{n}=H_{n}^{(0)}+M_{n} \label{defg} \end{equation} where $M_n$ is as before and $H_{n}^{(0)}$ is a matrix whose eigenvalue counting measure $N_{n}^{(0)}$ has a weak limit $N_{0}$ as $n\rightarrow \infty $. We denote by $f_{0}(z)$ the Stieltjes transform of $N_{0}$. In this case we use the resolvent formula (\ref{res}) for $B=H_{n}$ and $\ A=H_{n}^{(0)}$ and a natural extension of the above arguments. We obtain an analogue of Theorem 2.1 in which the Stieltjes transform $f(z)$ of the limiting counting measure $N$ is a unique solution of the functional equation \begin{equation} f(z)=f_{0}(z+2w^{2}f(z)) \label{defsc} \end{equation} belonging to the class (\ref{Nev}) . The equation defines the \textit{the deformed semicircle law} (see \cite{Kh-Pa:93} for its properties). \subsection{ Laguerre ensemble} The ensemble is defined as \begin{equation} M_{n}=\frac{1}{n}A_{n}A_{n}^{\ast } \label{Lag} \end{equation} where the $n\times n$ matrix $A_{n}$ has the probability distribution (cf.(\ref{gue})) \begin{equation} P(dA)=Z_{n}^{-1}\exp \left(-\frac{1}{2a^{2}}\hbox{ Tr } AA^{\ast }\right)dA \label{Lames} \end{equation} \begin{equation*} dA=\prod_{j,k=1}^{n}d\mbox{Re}A_{jk}\mbox{Im}A_{jk}. \end{equation*} In other words the entries $A_{jk},j,k=1,...,n$ of $A_{n}$ are independent complex Gaussian random variables defined by \begin{equation*} \mathbf{E}(A_{jk})=\mathbf{E}(A_{jk}^{2})=0,\quad \mathbf{E}% (|A_{jk}|^{2})=2a^{2}. \end{equation*} Note that the matrix $A$ is not Hermitian. The name of the ensemble is recent (see e.g. \cite{Pi:91}) and is related to the fact that in the orthogonal polynomial approach \cite{Me:91} one uses in the case of this ensemble the Laguerre polynomials (recall that in the case of the Gaussian Unitary Ensemble one uses the Hermite polynomials). The ensemble models a generic positively defined matrix. The real symmetric version of the ensemble in which $A$ is $n\times m$ matrix with statistically independent Gaussian entries is well known since the 30's in the multivariate analysis as the Wishart distribution and describes the sample covariance matrix of $m$ random Gaussian $n$-dimensional vectors \cite{Mu:82}. \begin{theorem} Let the Laguerre ensemble of random matrices be defined as above. Then its eigenvalue counting measure converges in probability 1 to the nonrandom measure of the form \begin{equation} N_{L}(\Delta )=\frac{1}{4\pi a^{2}}\int_{\Delta \cap \lbrack 0,8w^{2}]}\sqrt{% \frac{8a^{2}-\lambda }{\lambda }}d\lambda \label{IDSL} \end{equation} \end{theorem} \bigskip The proof of the theorem follows the scheme of that of Theorem 2.1, that is it is based on an analogue of Proposition 2.1. In particular, we have here an analogue of the important inequality (\ref{var}). However, the analogue of (\ref{m1}) has the form \begin{equation*} \mathbf{E(}g_{n}(z)\mathbf{)=-}\frac{1}{z}-2a^{2}\mathbf{E(}g_{n}^{2}\mathbf{% ).} \end{equation*} As a result, the corresponding quadratic equation is $2a^{2}zf^2+zf+1=0$ (cf.(\ref{qua})). This leads to (\ref{IDSL}). \bigskip \textbf{Remarks}. (i). Analogous results are valid for the cases when the matrix $A_{n}$ is real or quaternion. Thus we obtain the general formula for the density of the limiting measure \begin{equation} \rho _{L\beta }(\lambda )=\frac{1}{2\beta \pi a^{2}}\sqrt{\frac{4\beta a^{2}-\lambda }{\lambda },} \quad 0\leq \lambda \leq 4\beta a^{2}, \label{roL} \end{equation} that is concentrated on the interval $[0,4\beta a^{2}]$ for the orthogonal ($% \beta =1$), complex ($\beta =2$) and quaternion ($\beta =4$) cases respectively. (ii). We can consider an ensemble of more general form \begin{equation} H_{n}=H_{n}^{(0)}+M_{n} \label{defG} \end{equation} where $M_{n}$ is as in (\ref{Lag}) and $H_{n}^{(0)}$ is a matrix whose eigenvalue counting measure $N_{n}^{(0)}$ has a weak limit $N_{0}$ as $% n\rightarrow \infty $. We denote by $f_{0}(z)$ the Stieltjes transform of $% N_{0}$. In this case we use the resolvent formula (\ref{res}) for $B=H_{n}$ and $\ A=H_{n}^{(0)}$ and a natural extension of the above arguments. We obtain an analogue of (\ref{defsc})\ \ in which the Stieltjes transform $f(z) $ of the limiting counting measure $N$ is a unique solution of the functional equation \begin{equation} f(z)=f_{0}\left(z-\frac{2a^{2}}{1+2a^{2}f}\right) \label{Ladef} \end{equation} belonging to the class (\ref{Nev}). (iii) We can also consider a more general case when the random matrix $A_{n}$ is the $n\times m$ matrix with Gaussian i.i.d. entries. This case can be treated similarly. We discuss this case in more detail in Section 5 considering arbitrary distributed independent entries. \subsection{Circular Ensemble} The ensemble consists of $n\times n$ unitary matrices whose probability distribution is given by the normalized Haar measure on $U(n)$. The ensemble was introduced by Dyson together with its orthogonal and symplectic analogues (see \cite{Me:91} for references and results). We discuss below the simplest Circular Unitary Ensemble (CUE) but one can obtain similar results for two other ensembles. It is useful to write eigenvalues $\lambda _{j}$ of the ensemble in the form $\lambda _{j}=e^{i\theta _{j}},0\leq \theta _{j}<2\pi ,j=1,...,n$ and to introduce the normalized counting measure (cf.(\ref{N})) \begin{equation} N_{n}(\Delta )=\frac{1}{n}\natural \{\theta _{j}\in \Delta \} \label{NU} \end{equation} where $\Delta $ is Borel set of $[0,2\pi )$. An analogue of the Stieltjes transform (\ref{St}) for measures on the unit circle is the \textit{Herglotz transform} \cite{Ak-Gl:61} \begin{equation} h(z)=\int_{0}^{2\pi }\frac{e^{i\theta }+z}{e^{i\theta }-z}m(d\theta ),\quad |z|\neq 1. \label{Her} \end{equation} Respective inversion formula is (cf. (\ref{fp})) \begin{equation*} \int_{0}^{2\pi }\varphi (\theta )m(d\theta )=\lim_{r\rightarrow 1-0}\frac{1}{% 2\pi }\int_{0}^{2\pi }\varphi (\theta )\mbox{Re}h(re^{i\theta })d\theta . \end{equation*} We use here instead of (\ref{difg}) the differentiation formula \begin{equation} \mathbf{E}(\varphi ^{\prime }(M)AM)=\mathbf{E}(\varphi ^{\prime }(M)MA)=0 \label{difu} \end{equation} valid for any $C^{1}$ function $\varphi :U(n)\rightarrow \mathbf{C}$ and any Hermitian matrix $A$. This formula and the spectral theorem for unitary matrices according to which the Herglotz transform $h_{n}(z)$ of the eigenvalue counting measure can be written as (cf. (\ref{gG})) \begin{equation} h_{n}(z)=\frac{1}{n}\hbox{ Tr } \frac{U+z}{U-z} \label{hU} \end{equation} allow us to write the following relations for the moments of $h_{n}(z)$ : \begin{equation} \mathbf{E}(h_{n}(z))=-1,\quad |z|\neq 1, \label{mom1U} \end{equation} \begin{equation} \mathbf{E}(|h_{n}(z)|^{2})-|\mathbf{E}(h_{n}(z))|^{2}\leq \frac{C}{n^{2}}% ,\quad |z|\leq \frac{1}{4}. \label{varU} \end{equation} The first relation shows that $\mathbf{E}(N_{n}(\Delta ))=|\Delta|/ 2\pi$ for all $n$. This is easy to understand because the Haar measure is shift invariant. The second relation plays the role of (\ref{var}). By using these relations and following the scheme of proof of Theorem 2.1 one obtains \begin{theorem} Consider the ensemble of unitary matrices distributed according to the Haar measure on $U(n)$ (the CUE). Then the eigenvalue counting measure (\ref{NU}) of the ensemble converge in probability to the uniform measure on the unit circle. \end{theorem} \medskip \textbf{Remarks}. (i). Analogous statements are valid also for the Circular Orthogonal Ensemble and for the Circular Symplectic Ensemble (see \cite {Me:91} for their definitions and properties). (ii). Note that unlike Theorems 2.1 and 2.2 where we have the convergence with probability 1, Theorem 2.3 asserts only the convergence in probability, despite the bound (\ref{varU}). The reason is that in Theorems 2.1 and 2.2 we can consider $W_{n}$ and $A_{n}$ for all $n$ as defined on the same probability space of realizations of the semi-infinite matrices $W=\{W_{jk}\}_{j,k=1}^{\infty }$ and $A=\{A_{jk}\}_{j,k=1}^{\infty }$ equipped with the infinite product Gaussian measure. It is clear that similar natural and simple embedding does not exist for unitary matrices.\footnote {Although one can always use the probability space that is the product over all $n$ of the probability spaces consisting of the groups $U(n)$ with the normalized Haar measure as the probability measure.} This case can be regarded as an analogue of the triangular array scheme of probability and the Theorem 2.3 is an analogue of the (Tchebyshev) law of large numbers, while the case of the Gaussian and the Laguerre ensembles is analogous to the scheme of infinite number of i.i.d. random variables and Theorems 2.1 and 2.2 are analogues of the strong law of large numbers. To deduce the convergence in probability of $N_n(\Delta)$ for a fixed Borelien $\Delta$ from the convergence in probability its Herglotz transforms for a fixed $z, |z|\leq \frac{1}{4}$ one has to use the argument of Section 4 of \cite{Ma-Pa:67}. (iii). The simple differentiation formula (\ref{difu}) as well as its version (\ref{difUV}) below allows one to give a direct proof of the asymptotic freeness of unitary and diagonal matrices as $n\to \infty$ (see \cite{Vo-Di-Ni:92} for definitions and results and \cite{Xu:97, Ne-Sp:95, Vo:97} for some related recent results). Existing proofs are based on the representation of the Haar distributed unitary random matrices $U$ as the phase in the polar decomposition of the Gaussian distributed random matrix $X$ with complex i.i.d. entries and on the approximation of the phase by polynomials in $X$. Because of singularities of the polar decomposition representation $U=X(XX^{*})^{-1/2}$ these proofs are not simple to implement in all details. The approach based on the formula (\ref{difUV}), i.e. on the shift invariance of the Haar measure, seems more direct and simple (see \cite{Pa-Va:98} and Remark (iv) of Section 4). \section{Invariant Ensembles} In this Section we discuss the random matrix ensembles defined by the probability law \begin{equation} P(dM)=Z_{n}^{-1}\exp (-n \hbox{ Tr } V(M))dM \label{IE} \end{equation} where $M$ is a real symmetric or a Hermitian or a quaternion self-dual Hermitian matrix and $V$ is a bounded below and growing sufficiently fast at infinity function of respective matrix. For $V(\lambda)={\lambda^{2}}/{4w^{2}}$ we obtain the Gaussian Ensembles that were considered in the previous Section. In this paper we restrict ourselves by polynomial $V$'s. As it was in the case of the Gaussian Ensembles we discuss here the technically simplest Hermitian matrices. This subclass of ensembles (\ref{IE}) is motivated by Quantum Field Theory (see e.g. review \cite{DiF:95}). Following Quantum Field Theory we will call $V$ the potential. We will give below a result for convex $V$'s. More general case see in \cite{Kh-Pa-St:99}. \begin{theorem} Consider the random matrix ensemble consisting of Hermitian $n\times n$ matrices distributed according to (\ref{IE}) in which the potential $V$ is a convex polynomial of an even degree $2p$ and growing at infinity. Then the eigenvalue counting measure of the ensemble converges in probability \footnote{See Remark (ii) after Theorem 2.3.} to the nonrandom measure whose density is concentrated on the interval $(a,b)$ and has the form \begin{equation} \rho (\lambda )=p_{2p-2}(\lambda )\sqrt{(b-\lambda )(\lambda -a),}\quad a\leq \lambda \leq b \label{ro} \end{equation} where \begin{equation} p_{2p-2}(\lambda )=\frac{1}{\pi ^{2}}\int_{a}^{b}\frac{V^{\prime } (\lambda )-V^{\prime }(\mu )}{% \lambda -\mu }\frac{d\mu }{\sqrt{(b-\mu )(\mu -a)}} \label{p} \end{equation} is a positive on $(a,b)$ polynomial of the degree $2p-2$ and $a$ and $b$ are uniquely defined by the equations \begin{equation} \int_{a}^{b}\frac{\mu ^{q}V^{\prime } (\mu )d\mu }{\sqrt{(b-\mu )(\mu -a)}}=2\pi \delta _{1q},\quad q=0,1. \label{cond} \end{equation} \end{theorem} \bigskip The proof of the theorem follows again the scheme of the proof of Theorem 2.1. In particular we have the analogue of Proposition 2.1 \begin{proposition} Under the conditions of the preceding theorem the Stieltjes transform $% g_{n}(z)$ of the eigenvalue counting measure of the ensemble (\ref{IE}) has the following properties for $|\mbox{Im}z| \geq y$ and a certain $y$ depending on $V$ \begin{equation} \lim_{n\rightarrow \infty }\mathbf{E(}g(z))=f(z), \label{fV} \end{equation} \begin{equation} \mathbf{E(|}g_{n}^{2}(z)|)-|\mathbf{E(}g_{n}(z))|\leq \frac{const}{n^{2}} \label{vargV} \end{equation} where $f(z)$ is a unique solution of the quadratic equation \begin{equation} f^{2}+V^{\prime }(z)f+Q(z)=0 \label{quaV} \end{equation} satisfying (\ref{Nev}) and \begin{equation} Q(z)=\int \frac{V^{\prime }(z)-V^{\prime }(\mu )}{z-\mu }N(d\mu ) \label{Q} \end{equation} in which $N(d\mu )$ is the measure corresponding to $f$. \end{proposition} To prove the proposition we use the differentiation formula \cite {Be-It-Zu:80} \begin{equation} \mathbf{E}(\varphi^{\prime }(M)\cdot B)-n\mathbf{E}(\varphi (M)\hbox{ Tr } V^{\prime }(M)B)=0 \label{difV} \end{equation} valid for the matrix distribution (\ref{IE}), any function $\varphi:\mathbf {R} \to \mathbf {C}$ whose derivative is polynomially bounded on the whole real line and any Hermitian matrix $B$% . By applying this formula to the matrix element of the resolvent $% G(z)=(M-z)^{-1}$ we obtain the relation \begin{equation} \mathbf{E}(g_{n}^{2}(z))+\mathbf{E(}\frac{1}{n}\hbox{ Tr } G(z)V^{\prime } (M))=0 \label{g2V} \end{equation} where $g_n(z)$ is defined in (\ref{gG}). The identity \begin{equation} G(z)V^{\prime }(M)=G(z)V^{\prime }(z)+G(z)(V^{\prime }(M)-V^{\prime }(z)) \label{subst} \end{equation} allows us to rewrite this relation in the form \begin{equation} \mathbf{E(}g_{n}^{2}(z))+V^{\prime }(z)\mathbf{E(}g_{n}(z))+Q_{n}(z)=0 \label{prequa} \end{equation} where $Q_n(z)$ is defined as \begin{equation} Q_{n}(z)=\mathbf{E(}\frac{1}{n}\hbox{ Tr } Q_{n}(z,M)), \quad Q_{n}(z,M)= G(z)(V^{\prime }(M)-V^{\prime }(z\mathbf{))} \label{Q_n} \end{equation} and is a polynomial of the degree $2p-2$ if $V(z)$ is a polynomial of the degree $2p$. It is easy to see that for $V(z)=z^{2}/4w^{2}$ (\ref{prequa}) coincides with (\ref{qua}). Thus (\ref{prequa}) is an extension of (\ref{qua}% ) to the more general case of distribution (\ref{IE}), where $V(\lambda )$ is a polynomial of an even degree bigger than 2. An analogue of (\ref{prevar}) has the form \begin{equation} (1-\frac{2}{|\mbox{Im}z \cdot V^{\prime }(z)|})\mathbf{E(|}\gamma _{n}^{2}(z)|)\leq \frac{1}{|V^{\prime }(z)|}\mathbf{E(}\frac{1}{n}\hbox{ Tr } \gamma _{n}(z)q_{n}(z,M))+% \frac{1}{n^{2}|(\mbox{Im}z)^{3}V^{\prime }(z)|} \label{1} \end{equation} where \begin{equation*} q_{n}(z,M)=Q_{n}(z,M)-\mathbf{E(}Q_{n}(z,M)). \end{equation*} In the Gaussian case the first term in the r.h.s. of (\ref{1}) is absent. In view of the inequality \begin{equation*} |\mathbf{E(}\frac{1}{n}\hbox{ Tr } \gamma _{n}(z)q_{n}(z,M))|\leq \mathbf{E(}|\gamma _{n}(z)|^{2})^{1/2}\mathbf{E(}\frac{1}{n} \hbox{ Tr } q_{n}(z,M)q_{n}^{\ast }(z,M))^{1/2} \end{equation*} where $ \gamma _{n}(z)$ is defined in (\ref{gam}), it seems that the most natural way to obtain (\ref{vargV}) is to prove the estimate \begin{equation*} \mathbf{E(}\frac{1}{n} \hbox{ Tr } q_{n}(z,M)q_{n}^{\ast }(z,M))\leq \frac{const}{n^{2}} \end{equation*} thereby reducing the estimation of the variance of $g_{n}(z)$ to that of $% q_{n}(z,M)$. Unfortunately, we do not know the proof of the last estimate based on the differentiation formula (\ref{difV}) and similar to that in the second proof of Proposition 2.1. Thus we refer the reader to works \cite {ABM-Pa-Sh:95},\cite{Pa-Sh:97}, where the bound (\ref{vargV}) is proven for all $\mbox{Im}z|>0$ by using a combination of the orthogonal polynomials and variational techniques. A simple proof of a weaker version of (\ref{vargV}) with $n$ instead $n^{2}$ in the r.h.s. will be given in \cite{Kh-Pa-St:99} also by using the orthogonal polynomial technique. Any of these bounds allows us to replace $\mathbf{E(}g_{n}(z)^{2})$ in (\ref{fV}) by $\mathbf{E(}g_{n}(z)^{2})=\mathbf{E(}g_{n}(z))^{2}\equiv f_{n}^{2}(z)$. Besides, by applying (\ref{difV}) to $\varphi (M)=M$, we obtain the equality \begin{equation} \mathbf{E}(\frac{1}{n}\hbox{ Tr } MV^{\prime }(M))=1 \label{MV} \end{equation} that allows us to prove that all coefficients of the polynomial $Q_{n}(z)$ are uniformly bounded in $n$ . After that simple compactness arguments yield that the limit of any convergent subsequence $f_{n_{j}}(z)$ satisfies (\ref {quaV}). Having Proposition 3.1 we can prove Theorem 3.1 by the following arguments. By solving the quadratic equation (\ref{quaV}) we find that the measure $N$ has the bounded H{{\"o}}lder density $\rho $, that for a convex $V$ the support of the measure $N$ corresponding to $f$ is a finite interval $(a,b)$ and that (see \cite{Kh-Pa-St:99}) \begin{equation} v.p.\int_{a}^{b}\frac{\rho (\mu )d\mu }{\mu -\lambda }=-\frac{V^{\prime }(\lambda )}{2}, \quad \lambda \in (a,b) \label{seq} \end{equation} where the symbol $v.p.\int $ denotes the singular Cauchy integral. Regarding this relation as a singular integral equation for $\rho (\lambda)$ and using standard facts of the theory of singular integral equations \cite{Mu:53} \ we find that the bounded solution of the equation has the form \begin{equation} \rho (\lambda )=\frac{1}{\pi ^{2}}\sqrt{R(\lambda )} \int_{a}^{b}\frac{% V^{\prime }(\mu )-V^{\prime }(\mu )}{\mu -\lambda }\frac{d\mu }{\sqrt{R(\mu )% }},\quad R(\lambda )=(b-\lambda )(\lambda -a), \label{roV} \end{equation} provided that \begin{equation} \int_{a}^{b}\frac{V^{\prime }(\mu )d\mu }{\sqrt{R(\mu )}}=0. \label{solub} \end{equation} This gives condition (\ref{cond}) for $q=0$. Besides we have the normalization condition \begin{equation} \int_{a}^{b}\rho (\mu )d\mu =1 \label{normal} \end{equation} that can be rewritten in the form (\ref{cond}) for $q=1$ by using (\ref{roV}% ). It is clear that the integral is positive for a convex $V$ and that it is a polynomial of degree $2p-2$, if $V$ is a polynomial of the degree $2p $. The unique solubility of system (\ref{quaV}) can be proved by using the implicit function theorem \cite{Kh-Pa-St:99}. \textbf{Remark}. Consider the case of the monomial $V(\lambda)= |\lambda|^{2p}/2p$. In this case above formulas can be written in the form \begin{equation} \rho(\lambda)=\frac{1}{2\pi I_{2p-1}} \int^a_{|\lambda|}\frac{t^{2p-1}dt} {\sqrt{t^2-\lambda^2}}, \quad a^{2p}=\frac{\pi}{I_{2p}} \label{alpha} \end{equation} where $$ I_{\alpha}=\int^1_{0}\frac{t^{\alpha}dt} {\sqrt{1-t^2}} $$ These formulas are also valid for non-integer $p$, i.e. for potentials of the form $V(\lambda)= |\lambda|^{\alpha}/\alpha$ provided that $\alpha \ge 2$. For this case the formulas were obtained in \cite{Pa:92} by another method. They can also be obtained by a version of the method presented in this Section. In this version we use the identity \begin{equation} G(z)V^{\prime}(M)=G(z)V^{\prime }(\lambda)+G(z)(V^{\prime }(M)-V^{\prime }(\lambda)), \quad z=\lambda+i\varepsilon \label{subst1} \end{equation} instead of (\ref{subst}). It can be shown that this allows us to obtain the final formulas (\ref{ro})-(\ref{cond}) for example for non-polynomial (and even non-analytic) $V$'s provided that they are convex, even, grow faster than logarithmically at infinity and are of the class $C^{2}$ on any finite interval $(-L,L)$. \section{Law of Addition of Random Matrices} Consider Hermitian matrices of the form \begin{equation} V_{n}A_{n}V_{n}^{\ast }+U_{n}B_{n}U_{n}^{\ast } \label{UV} \end{equation} where $V_{n}$ and $U_{n}$ are random independent unitary matrices distributed both according to the normalized Haar measure on $U(n)$, $\ A_{n}$ and $% B_{n}$ are Hermitian matrices such that their normalized eigenvalue counting measures $N_{A_{n}}$ and $N_{B_{n}}$ converge weakly to the limits $N_{A}$ and $N_{B}$ respectively and satisfy the condition \begin{equation} \sup_n \int |\lambda |N_{A_n,B_n}(d\lambda )<\infty. \label{sup} \end{equation} \begin{theorem} The normalized eigenvalue counting measure of the ensemble of random matrices defined above tends in probability as $n\rightarrow \infty $ to the nonrandom limit whose Stieltjes transform $f(z)$ is a unique solution of the system of functional equations \begin{eqnarray} f(z) &=&f_{A}(z-\Delta _{B}(z)f^{-1}(z)) \notag \\ f(z) &=&f_{B}(z-\Delta _{A}(z)f^{-1}(z)) \label{sysad} \\ zf(z) &=&\Delta _{A}(z)+\Delta _{B}(z)-1 \notag \end{eqnarray} where $f$ belongs to the class (\ref{Nev}) and $\Delta _{A}$ and $\Delta _{B} $ are analytic for non-real $z$ and such that \begin{equation} \sup_{y\geq 1}y|\Delta _{A}(iy)|<\infty ,\quad \sup_{y\geq 1}y|\Delta _{B}(iy)|<\infty \label{delty} \end{equation} \end{theorem} \bigskip The theorem was proved in \cite{Sp:93} for the case of uniformly bounded in $% n$ matrices $A_{n}$ and $B_{n}$ by computing asymptotic form of moments of the sum via the moments of summands. This requires rather involved combinatorial analysis and impose the boundedness condition on matrices. We outline now the proof \cite{Pa-Va:98} based on the same ideas as above, i.e. on the resolvent identity and on a certain differentiation formula. The formula used in this case is \begin{equation} \mathbf{E}(\varphi ^{\prime } (UBU^{\ast })[UBU^{\ast },C])=0 \label{difUV} \end{equation} where $\varphi :\mathbf{R}\rightarrow \mathbf{C}$ is a $C^{1}$ function whose derivative is polynomially bounded on the real line, $B$ and $C$ are Hermitian matrices, $[B,C]=BC-CB$ and the symbol $\mathbf{E(...)} $ denotes the integration over $U(n)$ with respect to the Haar measure normalized to 1. The formula can be easily derived from the shift invariance of the Haar measure. Assume first that the norms of the matrices $A_{n}$ and $B_{n}$ are bounded uniformly in $n$. By applying (\ref{difUV}) to the resolvent identity (\ref{res}) relating the resolvent $G$ of matrix (\ref{U}) and the resolvent $G_{1}$ of matrix $A_{n}$ we obtain the matrix identity \begin{equation*} \mathbf{E}(G\frac{1}{n}\hbox{ Tr } G)=G_1\mathbf{E}(\frac{1}{n}\hbox{ Tr } G)-G_1\mathbf{E} (G \frac{1}{n}\hbox{ Tr } GUBU^{\ast }). \end{equation*} Assuming now asymptotic vanishing of the fluctuations of normalized traces, that we had in all cases above, we can rewrite this matrix identity in the form \begin{equation} \mathbf{E}(G)=G_{1}\left( z-\frac{\Delta _{B_{n}}(z)}{f_{n}(z)}\right) +o(1), \quad n\rightarrow \infty \label{G1} \end{equation} where \begin{equation} \Delta _{B_{n}}=\mathbf{E(}\frac{1}{n} \hbox{ Tr } U_{n}B_{n}U_{n}^{\ast }G), \quad f_{n}% =\mathbf{E}(\frac{1}{n} \hbox{ Tr } G). \label{Delta} \end{equation} and $|\mbox{Im}z|$ is large enough to guarantee inversibility of the argument of $G_{1}$ uniformly in $n$. Applying to (\ref{G1}) the operation $% n^{-1}\hbox{ Tr } $ we obtain the prelimit form of the first equation of system (\ref{sysad}). The second equation follows from the analogous procedure in which the roles of $A_{n}$ and $B_{n}$ are interchanged (recall that $U_{n}$ and $U_{n}^{\ast }$ have the same distribution). The third equation is the limiting form of the identity $n^{-1} \hbox{ Tr } G(z-A_{n}-U_{n}B_{n}U_{n}^{\ast })=1$ and of (\ref{Delta}). The unique solubility of (\ref{sysad}) follows from the implicit function theorem applicable for large $|\mbox{Im}z|$. The proof of the vanishing of the correlations, more precisely, a bound similar to (\ref{var}), is based on the same idea (see \cite{Pa-Va:98}). To obtain the general case (\ref{sup}) we truncate eigenvalues of $A_{n}$ and $B_{n}$ by a large number $T$ and use the minimax principle to control this procedure as $T\rightarrow \infty ,$ the compactness arguments and the unique solubility of (\ref{sysad}) in the class (\ref{Nev}), (\ref{delty}). \bigskip \textbf{Remarks}. (i). Since the normalized eigenvalue counting measure is unitary invariant \ and the Haar measure is shift invariant we can restrict ourselves without loss of generality to matrices of the simpler form \begin{equation} H_{n}=A_{n}+U_{n}B_{n}U_{n}^{\ast } \label{U} \end{equation} This form have, for example, the matrices (\ref{defg}) of the deformed GUE. Indeed, any matrix belonging to the GUE can be written in the form $\Psi _{n}\Lambda _{n}\Psi _{n}^{\ast },$ where $\Psi _{n}$ is the matrix of its eigenvectors, distributed uniformly over the $U(n)$ according to the Haar measure, $\Lambda _{n}$ is the random diagonal matrix of eigenvalues and $% \Psi _{n}$ and $\Lambda _{n}$\ are independent \cite{Me:91}. Besides, according to Theorem 2.1 the normalized eigenvalue counting measure of $% \Lambda _{n}$ converges with probability 1 to the semicircle law (\ref{sc}). Thus $\Psi _{n}$ plays the role of $U_{n}$, $\Lambda _{n}$ plays the role of $B_{n}$ and $N_{B}$ is given by (\ref{sc}). It can be easily checked that in this case the system (\ref{sysad}) reduces to (\ref{defsc}). Analogous fact is also valid for the deformed Laguerre ensemble (\ref{defG}) and also for certain classes of random operators acting in $l^2({\bf Z}^d)$ \cite{Ne-Sp:95}. (ii). It can be shown \cite{Pa-Va:98} that the theorem is also valid in the case when matrices $A_{n}$ and $B_{n}$ in (\ref{UV}) are also random, but independent of $U_{n}$ and $V_{n}$ and $N_{A_{n}}$ and $N_{B_{n}}$ converge weakly in probability to the nonrandom $N_{A}$ and $N_{B}$. Then the ensemble of deformed covariance matrices (\ref{defsc}) in which the random vectors $a_{l}$ are uniformly distributed over the unit sphere in $\mathbf{C}^{n}$ also has form (\ref{U}). As for the form (\ref{UV}), it is the case for the sum of two independent matrices distributed each according to the law (\ref{IE}) with possibly different polynomials $V_{1,2}$. In this case the condition (\ref {sup}) follows from (\ref{MV}). This case was considered in \cite{Ze:96} by using formal perturbation theory around the Gaussian ensemble. Thus, we see that Theorem 4.1 describes in a rather general setting the result of deformation of a random matrix by another matrix randomly rotated with respect to the first and allows us to find the limiting eigenvalue counting measure of the the sum (of the deformed or of the perturbed ensemble) provided that we know these measure for the both terms of the sum. (iii). For any function $f(z)$ satisfying (\ref{Nev}) one can introduce the ``selfenergy'' $\Sigma (z)$ by the relation \begin{equation} f(z)=-(z+ \Sigma (z))^{-1}. \label{self} \end{equation} It can be shown that $\Sigma (z)$ is also analytic for non-real $z$, and has the same property $\mbox{Im} \Sigma (z) \mbox{Im}z >0, \quad \mbox{Im}z \neq 0$ as $f$ (see (\ref{Nev})). Denote by $z(f)$ the functional inverse of $f(z)$ and set $\Sigma (z)=R(f)$. Then it is easy to see that (\ref{sysad}) is equivalent to the relation \begin{equation} R(f)=R_A (f)+R_B(f) \label{add} \end{equation} where $R_A (f)$ and $R_B(f)$ are the selfenergies corresponding to $f_A(z)$ and $f_B(z)$. The inverses of the Stieltjes transforms of limiting eigenvalue counting measures were used in \cite{Ma-Pa:67} in the qualitative study of the support of of these measures in the case (\ref{fMP}) below where \begin{equation} R(f)= - c\int \frac{t\sigma (dt)}{1+tf}. \end{equation} Relation (\ref{add}) was noted in \cite{Pa:72} for the case when $N_{A}$ and $N_{B}$ are both the semicircle laws (\ref{sc}), when $R(f)=w^{2}f$. The general form of this relation was proposed by D.Voiculescu in the context of the operator-algebras theory and its new branch known as the free probability theory (see \cite{Vo-Di-Ni:92} for results and references). In this theory the semicircle law plays the role of the Gaussian distribution and the measure defined by (\ref{roL}) (more generally, by formula (\ref {fMP}) below) plays the role of the Poisson distribution. (iv). Similar technique can be applied to multiplicative families of positive defined Hermitian and or unitary matrices and gives results \cite{Va:99} that generalize and simplify those of \cite{Ma-Pa:67} and also gives a more direct proof of certain results obtained for these ensembles in the context of free probability \cite{Vo-Di-Ni:92}. \section{Matrices With Independent and Weakly Dependent Entries. Tiny Perturbations} \subsection{Wigner Ensemble} The proofs outlined in previous Sections for simplest archetypal ensembles, the GUE first of all, can be elaborated and used in rather general case of Hermitian, real symmetric or self-dual Hermitian random matrices whose entries are independent or weakly dependent modulo symmetry conditions. The matrices can be written in the form (cf.(\ref{w})) \begin{equation} M_{n}=n^{-1/2}W_{n} \label{Wi} \end{equation} where matrix elements $W_{jk}^{(n)}$ of the matrix $W_{n}$ still satisfy (% \ref{mom}) but their probability laws $P_{jk}^{(n)}(dW)$ are not necessary Gaussian and may be $n$-dependent. We call these ensembles the Wigner Ensembles. In this general case we have to use instead of differentiation formula (\ref{difg}) the formula \cite{Kh-Kh-Pa:96} \begin{equation} \mathbf{E}(\xi \varphi (\xi ))=\sum_{a=1}^{s}\frac{\kappa _{a+1}\mathbf{\ \ }% }{a!}\mathbf{E}(\varphi ^{(a)}(\xi ))+\varepsilon _{s} \label{difs} \end{equation} where $\kappa _{a}$ are semi-invariants (cumulants) of a real-valued random variable $\xi $, $\varphi :\mathbf{R\rightarrow C}$ is a function of the class $C^{s+1}$ and $|\varepsilon _{s}|\leq C_{s}\sup_{x}|\varphi ^{(s+1)}(x)|% \mathbf{E}(|\xi |^{s+1})$. Another version of the method is based on the perturbation expansion of matrix elements of the resolvent in a particular matrix element of the matrix $M_{n}$. Indeed, according to Section 2 an important moment of the method is the asymptotical computation of the expectation \begin{equation} \mathbf{E}(\frac{1}{\sqrt{n}}\sum_{k=1}^{n}G_{jk}W_{kl}^{(n)}) \label{sum} \end{equation} basing on various differentiation formulas (see formulas (\ref{difg}),(\ref {difu}),(\ref{difV}), and(\ref{difs}) above). However, in the case of independent entries satisfying (\ref{mom}), this requires the knowledge of dependence of $G_{jk}$ on $W_{kl}^{(n)}$ up to linear terms only. Indeed, writing the resolvent identity (\ref{res}) for $B=n^{-1/2}W_n$ and $A=n^{-1/2}W|_{W_{kl}^{(n)}=0}$ we obtain \begin{equation} G_{jk}=G_{jk}^{kl}-n^{-1/2}(W_{kl}^{(n)}G_{jk}^{kl}G_{lk}^{kl}+ \overline{W_{kl}^{(n)}} G_{jl}^{kl}G_{kk}^{kl})+r_{n} \label{pert} \end{equation} where \begin{equation} G^{kl}=G|_{W_{kl}^{(n)}=0},\quad |r_{n}|\leq \frac{|W_{kl}^ {(n)}|^{2}}{n|\mbox{Im}z|^{3}}. \label{pert1} \end{equation} Substituting (\ref{pert}) in (\ref{sum}) and taking into account that $G^{kl}$ is independent of $W_{kl}^{(n)}$ we can perform explicitly the expectation with respect to $W_{kl}^{(n)}$ and obtain the relation \begin{equation} \mathbf{E}(g_{n}(z))=-1/z-\frac{w^{2}}{zn^{2}}\mathbf{E}( \sum_{k=1}^{n}G_{jl}^{kl}G_{kk}^{kl})+O(\max_{0\leq j,k\leq n}\mathbf{E}(|W_{jk}^{(n)}|^{3})/n^{1/2}|\mbox{Im}z|^{4}). \label{pert2} \end{equation} Now we can use (\ref{pert}) in the opposite direction to replace matrix element of $G^{kl}$ by those of $G.$ Thus, if \begin{equation} \sup_{n}\max_{0\leq j,k\leq n}\mathbf{E(}|W_{jk}^{(n)}|^{3})\le w_3<\infty, \label{W3} \end{equation} we obtain the analogue of (\ref{m1}) in the case of independent entries satisfying (\ref{mom}) and (\ref{W3}) with the error of the order $n^{-1/2}$. Similar arguments allows one to prove an analogue of (\ref{var}) with the r.h.s. of the order $n^{-1/2}$. This is sufficient for the proof of an analogue of Theorem 2.1 for the independent entries satisfying (\ref{mom}) and (\ref{W3}) and with convergence in probability instead of convergence with probability 1 (see also Remark (ii) after Theorem 2.3). \bigskip We list now several recent results obtained by combinations of approaches based on formulas (\ref{difs}) and (\ref{pert}) (for an account of previous results see \cite{Pa:96}). (i) \textbf{Semicircle Law.} The normalized eigenvalue counting measure converges weakly in probability to the semicircle law (\ref{sc}) if and only if in addition to (\ref{mom}) for any $\tau >0$ matrix elements (\ref{Wi}) of satisfy the condition \begin{equation} \lim_{n\rightarrow \infty }\frac{1}{n^{2}}\sum_{1\leq j\leq k\leq n}^{n}\int_{|W|\geq \tau n^{1/2}}|z|^{2}P_{jk}^{(n)}(dW)=0, \label{Lin} \end{equation} reminiscent the well known Lindeberg condition of the validity of the central limit theorem. This fact is known since the seventies, see \cite{Pa:72} for the sufficiency of somewhat stronger version of (\ref{Lin}) and \cite{Gi:75,Gi:90} for the necessity and sufficiency of (\ref{Lin}). However these results were obtained by rather complicated method. In \cite{Kh-Pa-St:99} we give a simple proof based on the approach of this paper. (ii) \textbf{1/n expansion} \cite{Kh-Kh-Pa:96}. By using (\ref{difs}) and assuming that $\{W_{jk}^{(n)}\}$ are identically distributed (modulo symmetry conditions as usually) and have $s+1$ finite moments one can construct $1/n$ - expansion of moments (\ref{momp}) in powers \ of $% 1/n^{l},l\leq s$, with the error of the order $1/n^{s+1/2}$ provided that the complex spectral parameter $z$ verifies (\ref{semip}). We give here the two results for the real symmetric matrices $M_{jk}^{n}=n^{-1/2}(1+\delta_{i,j})W_{jk}^{n}, \ \mathbf{E}(W_{jk}^{(n)})=0, \ \mathbf{E}(W_{jk}^{2})=w^2$ and \begin{equation*} \sup_{n}\mathbf{E}(|W_{jk}^{(n)}|^{5})<\infty. \end{equation*} We have then the following asymptotic formulas: \begin{equation*} m_{1}(z)=f_{sc}(z) \left \{1+\frac{1}{n}\left [\frac{w^{2}f_{sc}^{2}(z)}{% (1-w^{2}f_{sc}^{2}(z))^{2}}+\frac{\sigma f_{sc}^{4}(z)}{1-w^{2}f_{sc}^{2}(z)}% \right ] \right \}+O(n^{-\frac{3}{2}}), \end{equation*} where $\sigma =\mathbf{E}(|W_{jk}^{(n)}|^{4})-3\mathbf{E} (|W_{jk}^{(n)}|^{2})$ is known as the excess of random variable $W_{jk}^{(n)}$ and is assumed to be independent of $n$, and $f_{sc}$ is defined in (\ref{fsc}); \begin{equation*} m_{2}(z_{1},z_{2})-m_{1}(z_{1})m_{1}(z_{2})=n^{-2}c(z_{1},z_{2})+O(n^{-\frac{% 5}{2}}) \end{equation*} where \begin{equation} c(z_{1},z_{2})=\frac{2w^{2}}{% (1-w^{2}f_{sc}^{2}(z_{1}))(1-w^{2}f_{sc}^{2}(z_{2}))} \left \{w^{2}\left [\frac{% f_{sc}(z_{1})-f_{sc}(z_{2})}{z_{1}-z_{2}} \right ]^{2}+\sigma f_{sc}^{3}(z_{1})f_{sc}^{3}(z_{2}) \right \}. \label{cov} \end{equation} (iii) \textbf{Central Limit Theorem} \cite{Kh-Kh-Pa:96}. Assume in addition to (\ref{mom}) that the forth moments of $W_{jk}^{(n)}$ exist and are independent of $j,k$ and $n$. Then for $z$ from the domain (\ref{semip}) the random function $g_{n}(z)- \mathbf{E}(g_{n}(z))$ converges in distribution to the Gaussian random function with zero mean and the covariance (\ref{cov}). \subsection{Sample Covariance Matrices} In this Subsection we consider an ensemble of random matrices, generalizing the Laguerre ensemble of Subsection 2.2 and its real symmetric version known as the Wishart Ensemble of the sample covariance matrices. \cite{Mu:82}. Respective matrices have the form \begin{equation} M_{m,n}=\frac{1}{n}A_{m,n}T_{m}A_{m,n}^{*} \label{MP0} \end{equation} where $A_{m,n}$ are $n\times m$ random matrices whose entries $A_{jk}^{(m,n)}$ are i.i.d. complex random variables satisfying conditions \begin{equation} \mathbf{E}(A_{jk}^{(m,n)})=\mathbf{E}((A_{jk}^{(m,n)})^{2})=0,\quad \mathbf{E% }(|A_{jk}^{(m,n)}|^{2})=1, \label{MP2} \end{equation} \begin{equation} \sup_{m,n}\max_{1\leq j\leq n,1\leq k\leq m}\mathbf{E}% (|A_{jk}^{(m,n)}|^{4}) \leq a_4 <\infty , \label{MP4} \end{equation} and $T_{n}$ is a diagonal matrix. We assume that \begin{equation} m\rightarrow \infty ,\quad n\rightarrow \infty ,\quad \frac{m}{n}\rightarrow c<\infty , \label{c} \end{equation} and that the normalized counting measure \begin{equation} \sigma _{m}(\Delta )=\frac{1}{n}\natural \{t_{l}\in \Delta \} \label{sigma_m} \end{equation} of eigenvalues $t_{l},l=1,...,m$ of $T_{n\text{ }}$ has a weak limit \begin{equation} \sigma _{m}(\Delta )\rightarrow \sigma (\Delta ),\text{ \ }m\rightarrow \infty . \label{sigma} \end{equation} In particular, $t_{l},l=1,...,m$ may be i.i.d. random variables independent of $A_{m,n}.$ \begin{theorem} Under conditions listed above the eigenvalue counting measure of matrices (\ref{MP0}) converges weakly in probability to the nonrandom measure whose Stieltjes transform is a unique solution of the functional equation \begin{equation} f(z)=- \left (z-c\int \frac{t\sigma (dt)}{1+tf(z)}\right )^{-1} \label{fMP} \end{equation} in the class (\ref{Nev}). \end{theorem} We outline the scheme of the proof, based on the same idea as above, i.e. on the careful analysis of the result of infinitesimal as $n\rightarrow \infty$ changes of respective matrices. Start again from the resolvent identity (\ref{res}) written for the pair $B=M_{n},$ $A=0.$ Applying to the identity the operation $\mathbf{E}(n^{-1} \hbox{ Tr }...)$ we obtain \begin{equation} \mathbf{E}(g_{n}(z))=-\frac{1}{z}-\frac{1}{zn^2}\sum_{l=1}^{m}t_{l}\mathbf{E} (\sum_{j,k=1}^{n}A_{jl}^{(m,n)}G_{jk}\overline{A_{kl}^{(m,n)}}) \label{g1} \end{equation} We can use now the scheme of proof of Theorem 2.1 using (\ref{difs}) instead (\ref{difg}). It is more convenient however to apply here a somewhat different scheme. It is analogous to that based on relations (\ref{sum}), (\ref{pert}% ), and (\ref{pert1}) in the case of the Wigner Ensembles of the previous subsection, however applied not to individual matrix elements but to the columns $a_{l}=\{n^{-1/2}A_{jl}^{(m,n)}\}_{j=1}^{n},l=1,...,m$ of the random matrix $A^{(m,n)} $. Treating the columns as vectors of $\mathbf{C}^{n}$ we can rewrite (\ref{g1}% ) as follows \begin{equation} \mathbf{E}(g_{n}(z))=-\frac{1}{z}-\frac{1}{zn}\sum_{l=1}^{m}t_{l}\mathbf{E} ((Ga_{l,}a_{l})) \label{g2} \end{equation} where $(.,.)$ is the scalar product in $\mathbf{C}^{n}.$ Since vectors $a_{l} $ are independent we perform first the asymptotic computation of the expectation with respect to $a_{l}$ in the $l$-th term of the sum like we did in the previous subsection for $W_{jk}^{(n)}$. To this end we use the formula, giving in the explicit form the result of perturbation of the resolvent $G_{C}$ of an arbitrary matrix $C$ by the rank one matrix $L_a,a\in \mathbf{C}^{n}$ defined by its action on any vector $x \in \mathbf{C}^{n}$ as $L_a x=(x,a)a$: \begin{equation} (C+L_a-z)^{-1}=G_{C}-G_{C}L_aG_{C}(1+(G_{C}a,a))^{-1}. \label{Kr} \end{equation} The formula can be easily derived from the general resolvent identity (\ref {res}). By applying the formula to $C=M_{n}|_{a_{l}=0}$ we obtain that \begin{equation*} (Ga_{l},a_{l})=-\frac{t_{l}(G_{l}a_{l},a_{l})}{1+t_{l}(G_{l}a_{l},a_{l})} \end{equation*} where $G_{l}=G|_{a_{l}=0}$. This relation will play the role of (\ref{pert}). Indeed, assume first that for some finite $T$ and $a_2$ \begin{equation} \sup_n \max_{1 \leq l \leq m} |t_l|\leq T, \quad \sup_n \max_ {1 \leq l \leq m}||a_l||\leq a_2 \label{T} \end{equation} Since $G_{l}$ does not depend on $a_{l}$ and since random vectors $\{a_{l}\}_{l=1}^{m}$ are mutually independent one can find from (\ref{MP2}) and (\ref{MP4}) that \begin{eqnarray*} \mathbf{E}_{l}((G_{l}a_{l},a_{l}))=\frac{1}{n}\hbox{ Tr } G_{l} \equiv g_n^{(l)} \\ \mathbf{|E}_{l}((G_{l}a_{l},a_{l}))-\frac{1}{n}\hbox{ Tr } G_{l}|^{2} \leq &\frac{Ca_4% }{n|\mbox{Im}z|^{2}} \end{eqnarray*} where the symbol $\mathbf{E}_{l}(...)$ denotes the operation of the expectation with respect the vector $a_{l}$ only. These relations allow us to present (\ref{g2}) in the form (cf. (\ref{pert})) \begin{equation} \mathbf{E}(g_{n})=-\frac{1}{z}-\sum_{l=1}^{m}t_{l}\mathbf{E} \left (\frac{g_{n}^{(l)}}{1+t_{l}g_{n}^{(l)}}\right)+r_{n} \label{g3} \end{equation} where now \begin{equation*} |r_{n}|\leq \frac{C(1+a_{4})}{n|\mbox{Im}z|^{2}},\text{ \ \ }|\mbox{Im}% z|\geq y_{0}, \end{equation*} and $y_{0}$ depends on $T$ and on $a_2$ of (\ref{T}). Besides, applying to (\ref{Kr}) the operation $\frac{1}{n}\hbox{ Tr }...$ we obtain that \begin{equation*} g_{n}-g_{n}^{(l)}=-\frac{1}{n} \cdot \frac{t_{l}(G_{l}^{2}a_{l},a_{l})}{% 1+t_{l}(G_{l}a_{l},a_{l})} \end{equation*} and thus \begin{equation*} |g_{n}(z)-g_{n}^{(l)}(z)|\leq \frac{1}{n|\mbox{Im}z|}. \end{equation*} By using three last relations we can write instead (\ref{g3}) for $|% \mbox {Im}z|\geq y_{0}$ and n$\rightarrow \infty $ (cf. (\ref{pert2})) \begin{equation} \mathbf{E}(g_{n}(z))=-\frac{1}{z}- \int \mathbf{E}\left(\frac{tg_{n}(z)}{1+tg_{n} (z)}\right)\sigma _{m}(dt)+o(1) \label{g4} \end{equation} where $\sigma _{m}(dt)$ is defined in (\ref{sigma_m}). This is an analogue of (\ref{m1}). Similar arguments allow us to prove an analogue of (\ref{var}% ). As a result we obtain (\ref{fMP}) in the case of bounded $t_{l}$ and $% a_{l}.$ General case can be obtained from the proven one by using the the analyticity of the Stieltjes transform up to the real axis, the truncation of $t_l$ and $a_l$, the minimax principle to control the truncation procedure, the compactness arguments, and the unique solubility of (\ref{fMP}) in the class (\ref{Nev}). The latter results from the the implicit function theorem. \bigskip \textbf{Remarks}. (i). In the case when $m=n$ and $\sigma (dt)$ has only one atom at $t=2a^2$ we obtain (\ref{Ladef}). (ii). Similar arguments shows that the eigenvalue counting measure of deformed ensemble (\ref{MP0}) \begin{equation} H_{n}=H_{n}^{(0)}+M_{m,n} \label{defMP} \end{equation} where $M_{m,n}$ is defined by (\ref{MP0}) and $H_n^{(0)}$ has the limiting eigenvalue counting measure $N_0$ (like in (\ref{defg}) and in (\ref{defG}) also tends weakly in probability to the nonrandom limit whose Stieltjes transform is a unique solution of the functional equation \begin{equation} f(z)=f_{0}\left (z-c\int \frac{t\sigma (dt)}{1+tf(z)}\right ) \label{genMP} \end{equation} This functional equation was derived first in \cite{Ma-Pa:67} by another and rather complicated method. The method was based on the study of the sequence of matrices $H_n^{(p)}, \quad p=1,...,m$ defined as \begin{equation} (H_n^{(p)})_{jk}=(H_n^{(0)})_{jk}+\sum_{l=1}^p t_l A^{(m,n)}_{lj} \overline {A^{(m,n)}_{lk}} \label{Hp} \end{equation} and ``interpolating'' between $H_n^{(0)}$ and $(H_n^{(n)})=H_n$. Asymptotic computation of the differences $\hbox{ Tr } (H_n^{(p+1)}-z)^{-1}- \hbox{ Tr } (H_n^{(p)}-z)^{-1}$ based on formula (\ref{Kr}) lead to the first order partial differential equation for \begin{equation*} f(t,z)=\lim_{n \to \infty, p/n \to t} \frac{1}{n}\hbox{ Tr } (H_n^{(p)}-z)^{-1}. \end{equation*} Solving the differential equation subject the conditions $f(0,z)=f_0(z), f(1,z)=f(z)$ one obtains (\ref{genMP}). The derivations of equation (\ref{genMP}) given later in \cite{Gi:75,Gi:90, Pa-Kh:89, Kh:96} and as well as the proof outlined above are more simple and direct. On the other hand, the sequence $H_n^{(p)}$ in (\ref{Hp}) can be regarded as a matrix version of the sum of independent random variables with varying upper limit used often in the study of limit theorems and processes with independent increments. Similar observation was used recently in \cite{Vo-Di-Ni:92} to construct free (non-commutative) analogues of these processes where, in particular, an analogous partial differential equation was obtained (called in \cite{Vo-Di-Ni:92} the complex Burgers equation). \section*{Acknowledgment} I am thankful to A.Khorunzhy, B.Khoruzhenko, A.Stoyanovich and V.Vasilchuk with whom the results mentioned above were obtained. The final version of this paper was written when I was participating the semester ``Random Matrices and their Applications'' at the MSRI. I am grateful to Prof.D.Eisenbud for the hospitality at the Institute, and to organizers of the semester P.Bleher and A.Its for the invitation. Research at MSRI is supported in part by NSF grant DMS-9701755.
\part{#1} \pagestyle{myheadings} \markright{Eaves, Ph.D\hfill1999\hfill\ Page } } \newcommand{\myRef}[1]{ \S\ref{#1} } \newcommand{\textit{etc.}}{\textit{etc.}} \newcommand{\textit{e.g.}\,}{\textit{e.g.}\,} \newcommand{\textit{i.e.}\,}{\textit{i.e.}\,} \newcommand{\textit{vs.}\,}{\textit{vs.}\,} \newcommand{\textit{et al.}\,}{\textit{et al.}\,} \newcommand{\textit{op. cit.}\,}{\textit{op. cit.}\,} \newcommand{\textit{viz.}\,}{\textit{viz.}\,} \newcommand{\textit{NB.}\,}{\textit{NB.}\,} \newcommand{\textit{vs.}\,}{\textit{vs.}\,} \newcommand{\emph{Aidan}\,}{\emph{Aidan}\,} \newcommand{\emph{CoBase}\,}{\emph{CoBase}\,} \newcommand{\emph{Java}\,}{\emph{Java}\,} \newcommand{\emph{Jigsaw}\,}{\emph{Jigsaw}\,} \newcommand{\emph{PostgreSQL}\,}{\emph{PostgreSQL}\,} \newcommand{\mySrc}[1]{\textsf{#1}} \newlength{\facewd} \newlength{\faceht} \newcommand{\markOver}[1]{ \settowidth{\facewd}{#1} \settoheight{\faceht}{#1} \raisebox{\faceht}[0pt]{ \makebox[0pt][l]{ \hspace{.15\facewd}$\blacktriangledown$ } } #1 } \newcommand{\markUnder}[1]{ \settowidth{\facewd}{#1} \settoheight{\faceht}{#1} \raisebox{-\faceht}[0pt]{ \makebox[0pt][l]{ \hspace{.15\facewd}$\blacktriangledown$ } } #1 } \newcommand{\myImport}[1]{ \markOver{\mySrc{#1}} } \newcommand{\myExport}[1]{ \markUnder{\mySrc{#1}} } \newcommand{\myAttr}[1]{ \mySrc{#1} } \newcommand{\myEnv}[1]{ \texttt{#1} } \newenvironment{myList}[1]{ \begin{list}{\#1{bean}}{\usecounter{bean}} \setlength{\rightmargin}{\leftmargin} } { \end{list} } \begin{document} \title{ODP channel objects that provide services transparently for distributing processing systems} \author{Walter D Eaves} \date{\today} \maketitle \begin{abstract} This paper describes an architecture for a distributing processing system that would allow remote procedure calls to invoke other services as messages are passed between clients and servers. It proposes that an additional class of data processing objects be located in the software communications channel. The objects in this channel would then be used to enforce protocols on client--server applications without any additional effort by the application programmers. For example, services such as key--management, time--stamping, sequencing and encryption can be implemented at different levels of the software communications stack to provide a complete authentication service. A distributing processing environment could be used to control broadband network data delivery. Architectures and invocation semantics are discussed, Example classes and interfaces for channel objects are given in the \emph{Java}\, programming language. \end{abstract} \section{Distributed Processing Platforms} \subsection{RPCs and Connections} A distributed processing platform provides a method of making a Remote Procedure Call, an RPC, almost transparent to the application developer---see, for example, \textit{Tivoli}\cite{orb:tivoli} and \textit{Orbix} \cite{orb:orbix}---the relatively mature industry standard for both of which is the Common Object Request Broker Architecture, \textit{CORBA}, from the Object Management Group\cite{web:corba}, OMG; this defines an Object Request Broker, ORB, which is an infrastructure for RPCs. Recently \textit{Sun Microsystems} \cite{web:sun} have enhanced the Remote Method Invocation package, \mySrc{java.rmi}, for \emph{Java}\,\cite{web:java} as part of the \emph{Java}\, Development Kit, JDK, 1.2. It now allows connections to be created between a client and a server which can have a different data transfer representation \cite{java:rmi:sockets}. As pointed out in the documentation for this feature, this is particularly suitable for implementing the Secure Socket Layer, SSL, \cite{draft:ssl} and could also be used to implement the proposed successor to SSL Transport Level Security, \cite{draft:tls}. \paragraph{Open Distributed Processing Architecture} A suitable architecture to exploit this new functionality in \mySrc{java.rmi} and in other distributed processing platforms has been proposed in the Open Distributed Processing standards, ISO--ODP, \cite{ISO:ODP}. \subparagraph{Bindings and Channels} The prescriptive model, \cite{ODP:presc}, generalizes the concept of connection between client and server to be a logical binding which is realized by both client and server using a channel, see also \cite{sft:ansaware}. Figure \ref{fig:odp-channel} illustrates the concepts of bindings and channels. Application objects have bound with one another probably using a naming service; they have negotiated a channel configuration that requires two objects: a data presentation conversion object and a data transport object. The channel configuration is fixed by bindings. The client sends data; the server receives. The data is passed from the application object to the channel which carries out the conversion and the transport to the server. \begin{figure}[htbp] \begin{center} \includegraphics[angle=-90, keepaspectratio=1, totalheight=6in]{odp-channel.eps} \caption{Bindings across two channels} \label{fig:odp-channel} \end{center} \end{figure} \begin{description} \item[Binding] A binding is a contract between the client and the server stating the parameters by which they will communicate. A binding between two transport objects would usually specify: \begin{itemize} \item Transport layer protocol to be used, \textit{e.g.}\, UDP/IP or TCP/IP or Pipe/Unix. \item Transport layer parameters \item Network addresses for chosen transport layer protocol \end{itemize} The presentation layer object would convert from the the local data representation to a network representation. The only specification needed here is the source and target representations. In this case, the binding is implicit in the implementation of the objects, they need not be initialized with parameters. More sophisticated systems will have higher demands and will insist that other services be used as well, for example: \begin{itemize} \item Data security \item Transaction management \item Call billing \item Data compression \item Relocation manager \end{itemize} These would all require that configuration parameters be specified and may also demand that they are operate together. \end{description} There is no need for channels to be symmetric, the server could implement a call--logging object in its channel without having an object of the same type in the client's channel. Bindings do need to be current. A transport object might close a connection, in which case, it would no longer be current, but if it were to leave enough information to allow a re--connection, then it is, in effect, still current. \subparagraph{Tri--partite Bindings} Bindings need not be bi--partite. An application service that would require a tri--partite binding is relocation management, illustrated in figure \ref{fig:odp-3binding}. The idea of which is that should the server choose to relocate, it would notify a relocation manager of its new addresses and move there. When the client calls the server at its old address and fails to reach it, the relocator object in the client's channel would call the third party and ask for the new address of the server, establish a new set of bindings with it, \textit{i.e.}\, construct a new channel and destroy the old, and send the message again. This would all be transparent to the application object. The relocation object in the channel would need to call the relocation manager for the new addresses and would thus become an application object; it would need to establish its own channel with the relocation manager. \begin{figure}[htbp] \begin{center} \includegraphics[keepaspectratio=1, totalheight=4in]{odp-3binding.eps} \caption{Relocation: a tri--partite binding} \label{fig:odp-3binding} \end{center} \end{figure} \subsection{Inflexible System Designs} \label{sec:inflexible} Although the ISO--ODP defines a flexible architecture, most implementors of distributed processing platforms provide application programmers with inflexible systems: the channel can only contain a presentation and a transport object. \paragraph{Presentation: Stubs and Skeletons} Both \textit{CORBA} ORBs and the RMI defined in \emph{Java}\, make use of ``stubs'' and ``skeletons'' and a stub--compiler. The term infrastructure will be used for the engineering that realizes an ORB or RMI. \begin{description} \item[Stubs] These are used by a client when invoking a method on a remote server. The client does not have a local implementation for the service the remote server has, but it will have a definition of its interface which can be used as if it were the implementation of the service. It can present the interface definition to a stub--compiler which will generate code to invoke each method on the remote server. The stub acts as a proxy for the remote server in the client's address space. The stub for each method needs to do the following things: \begin{enumerate} \item Construct a request A request object is a container entity that carries the invocation to the server; it contains: \begin{itemize} \item The address of the remote server \item The name of the method being invoked, possibly with version control information for the interface. \item The parameters for the method \item The return address \end{itemize} The stubs can also contain the code to \emph{marshall}\footnote{The verb is \textit{to marshal}, but this is so often mis-spelt that to marshall has become acceptable.} the parameters into a universal transfer presentation. In this form, the stubs also comprise the presentation object. \item Invoke the request The request, now just a sequence of bytes with an associated address for the server and a return address for the sender, is passed to the infrastructrure which sends it to the network socket for the server. \item Get the reply The server will return a reply in another container type. \item Re--construct the reply The reply is then re--constructed or, rather, its contents are \emph{un--marshalled} and returned to the client. Again, this may be part of the code in the stubs. \end{enumerate} \item[Infrastructure] It might be best now to explain how the infrastructure manages to provide the RPC service. \begin{enumerate} \item At the server on creation An application programmer defines an interface and produces an implementation for it. A program that effectively acts as a loader issues instructions to the infrastructure to create a socket to receive requests for that server and will associate the remote server implementation with that socket. \item At the naming service The application programmer will ensure that the address of his newly--created remote server is put into a well--known naming service. The client will then collect the address from the naming service. \item At the client on invocation The address collected at the naming service will contain enough information to allow the client's infrastructure to send the request container as a stream of bytes to the socket that the server's infrastructure has associated with the remote server's implementation. \item At the server on invocation The network socket will be activated by the client's infrastructure (a connect and a data send) and the server's infrastructure will collect the data at the socket and, because it has recorded the object responsible for that socket, it can activate the server skeleton. \end{enumerate} \item[Skeletons] \label{sec:inflexible:1} These are invoked by the server's infrastructure when the network socket for a server is activated and the data comprising a call has been collected. It invokes the implementation of the method the client wants to use at the server. A skeleton can consist of one method that unmarshalls enough of the request container to be able to look up the method to invoke---this is known as \emph{dispatching}. This partial unmarshalling has to be done in this way, because each method will unmarshall the remainder of the request container differently to obtain the parameters to pass to the server's method implementation. After the invocation is made the results will be marshalled into a reply container which is then sent back to the client. \end{description} The stubs and the skeleton effectively form the presentation channel object. Stubs have to be produced by a stub--compiler, but skeletons can be made generic, if the underlying infrastructure supports a reflective invocation mechanism, see \mySrc{java.lang.reflect} or the \textit{CORBA} Dynamic Invocation Interface. Some stubs generated by the \emph{Java}\, \mySrc{rmic}, Remote Method Interface Compiler, are given in appendix \ref{cha:stubs}. These demonstrate the use of reflective language features. \paragraph{Transport} When the request containers are passed to the infrastructure the transport object is eventually invoked. Most ORBs only provide one transport mechanism which sends data through TCP/IP sockets to its destination. Although some systems do allow different protocols: UDP/IP or Pipe/Unix. \paragraph{Limitations} The problem with this is that if one wants to implement any useful application services---encryption, billing and so forth---the infrastructure does not help. For example, to encrypt and decrypt data sent as part of a remote procedure call, one would have to implement one's own stubs and skeletons, see figure \ref{fig:ODP-channel-0}. \begin{figure}[htbp] \begin{center} \includegraphics{ODP-channel-0.eps} \caption{Encrypting and Decrypting by the application programmer} \label{fig:ODP-channel-0} \end{center} \end{figure} The application programmer has to construct a call, marshall the data, encrypt it, send it using a generic method, which will marshall it again. At the server, the data would be delivered to the generic method, and the application programmer would then have to decrypt it, unmarshall the decrypted data, reconstruct the call and dispatch it. After dispatching, collect the results, marshall, encrypt and return the reply. \subsection{More flexible: \emph{Java}\, RMI Custom Socket Factories} A more flexible implementation has been provided by \textit{Sun} in \emph{Java}\,. It allows a different type of socket to be used as the transport object. \paragraph{Method} The custom socket factories have to implemented in the following way: \begin{enumerate} \item Derive and implement classes for the new socket type's datastream from \mySrc{java.io.FilterOutputStream} and \mySrc{java.io.FilterInputStream}, call them \mySrc{MyOutputStream} and \mySrc{MyInputStream}. \item Derive and implement classes for the new socket types \mySrc{java.net.Socket} and \mySrc{java.net.ServerSocket} that use the new streams \mySrc{MyOutputStream} and \mySrc{MyInputStream}. \end{enumerate} Then create socket factory implementations that can be used by RMI. \begin{enumerate} \item A client-side socket factory that implements \mySrc{RMIClientSocketFactory} and implement the \mySrc{RMIClientSocketFactory.createSocket()} method. \item A server-side socket factory that implements \mySrc{RMIServerSocketFactory} and implement the \mySrc{RMIServerSocketFactory.createServerSocket()} method. \end{enumerate} Then one has to ensure that the constructor for the remote server is told to use the new socket factories. The infrastructure creates the new type of socket when demanded and invokes the create socket methods. The \mySrc{RMISecurityManager} at the client will then determine that a particular type of socket has to be used and will load the custom socket factory implementations. \paragraph{Possibilities: Implementing SSL} Using custom socket factories, it is possible to implement a Secure Socket Layer. The \mySrc{RMIClientSocketFactory.createSocket()} would be used to perform the key exchange with the server and the custom input and output streams would apply the session key to encrypt on send and decrypt on receive. \paragraph{Limitations} Unfortunately, using custom socket factories only increases the variety of the transport objects that can be employed, it does not allow different kinds of objects to be placed in the channel. \section{Channel Objects} What is needed is a means of placing objects in the channel before and after the presentation object. These objects should have a simpler instantiation and invocation procedure than using the custom socket method in \emph{Java}\,. \subsection{Some Requirements} \begin{enumerate} \item Different interests \- System Configurable The objects placed in the channel between client and server are the result of a negotiated agreement between the client, its server and their respective environments. It is well--known that security requirements for messages depend on the workstation that the client is using \cite{sec:sesame}, which may be connected to a secure local area network on which both the the client and server reside and so, for example, no security measures need be taken; or, the client could be accessing the server remotely from the Internet through a modem in which the server might require that the client use encryption. It would be desirable if the client and the server could both specify their requirements and some negotiation take place that could create a mutually acceptable protocol stack in the software communications channel. \item Different Methods for Different Actions A request and reply actually require that four different channels be traversed by messages, see figure \ref{fig:odp-4channels}: \begin{enumerate} \item Request \- this channel is created by the infrastructure for the client and sends a message over the network. \item Indication \- this channel is used to receive from the network and is created by the infrastructure for the server. \item Response \- created by the infrastructure for the server to return the results of the client's request message. \item Confirmation \- receives from the network and is created by the infrastructure for the client. \end{enumerate} Should an error occur at the server it is returned via the response and confirmation messages. The stubs are responsible for managing the thread of execution of the application object: once a message is put onto the request channel, the application object's thread can be suspended whilst it waits for the reply to arrive on the confirmation confirmation channel. If the message is a cast of some kind (broadcast, multi--cast or one--cast) there will be no response or confirmation. \begin{figure}[htbp] \begin{center} \includegraphics[angle=-90, keepaspectratio=1, totalheight=6 in]{odp-4channels.eps} \caption{Four channels are used in a request and response} \label{fig:odp-4channels} \end{center} \end{figure} \item Different Implementations Most channel objects are derived from the same source and will have the same implementation. \emph{Java}\, allows classes implementing different objects to be loaded over the network, so it would be possible for the client and server to agree upon and load the same class implementation, which they would be able to do with the custom socket factories method. This may not always be the case, some channel objects may be optimized to make use of a different operating system, but provide the same functionality. Audio and video data streaming are good examples of this need: some micro--processors now have support for stream data. \item Transparency \begin{enumerate} \item Management It would be desirable if the additional services provided by the channel objects did not need to be initialized or managed by either the client or the server application programs but they were activated by their respective infrastructures. \item Exception handling One of the difficulties of developing applications in enterprise environments is that as messaging becomes more sophisticated---supporting for example, confidentiality, authorization, call--billing---the number of possible errors increases because each of these sub--systems introduces new ones. It must be possible for channel objects to clear down their own errors, so that errors returned to application objects only involve the application. \item Using Possession rather than Inheritance for Coupling It would also be desirable if the classes for the channel objects did not extend the existing class hierachies of the message transmission sub--system, in the way that custom socket and socket factories do. Extending class hierachies is not as flexible as specifying an order of invocation. \item Interface Definitions Suitable for Reflection. It would also be desirable if the infrastructures could load channel objects' classes remotely and have a simple enough invocation syntax so that reflection mechanisms could be used to invoke the channel objects \textit{without requiring a stub compiler}. \emph{Java}\, already does this with its version 1.2 stubs, using package \mySrc{java.lang.reflect}, and \textit{CORBA} has a Dynamic Invocation Interface which can achieve the same goal. \end{enumerate} \item Efficiency It would also be desirable if the channel objects did not create a large stack of calls; primarily because some target environments for remote procedure call platforms will be embedded systems on SmartCards \cite{smartcard:tech}. \end{enumerate} \subsection{Some Nomenclature} With the aid of figure \ref{fig:odp-4channels}, it is possible to be more precise with the terms used: \begin{itemize} \item Initiator: starts a four--phase call sequence; may also be called the requestor. \item Acceptor: accepts the call made upon it; may also be called the responder, because it generates the response. \end{itemize} Both an initiator and an acceptor will send and receive as part of four--phase call sequence. Ordinarily, the client will be the initiator of all calls, but some remote procedure call systems permit call--backs, in which case the server is the initiator and the client the acceptor. \section{Architectures for Channel Objects} There are basically two types of architecture that could support channel objects. \subsection{Stream--oriented Architecture} The custom socket factory architecture is stream--oriented. It acts upon the data being sent between client and server as a stream applying a data transformation to the data when it is sent and undoing this transformation at the other end. The data is treated as opaque and can be delivered in packets as small as one byte. The implementor of the stream handler does not know whether the data has just started or is about to finish. One of the attractions of this approach is that many of the operations that data networks perform on data can be implemented in software: segmenting, and its converse re--assembling, can be implemented easily and this would allow remote procedure call systems to make use of packet--oriented transmission, such as UDP/IP. Segmenting and re--assembly could be implemented as channel objects and data would be segmented and then sent as packets on a UDP socket for re--assembly by another channel object at the server. Streams can be chained---the output of one providing the input for the next. Most operating systems support \textit{pipes} to do this: \emph{Java}\, supports the \mySrc{java.io.PipedInputStream}, and output, classes. Multi--casting could be easily achieved with pipes: simply have a pipe that sends on a socket and echoes its input; then connect a series of these together. A stream--oriented architecture is better for real--time data delivery. All of the processing in the stream--handler is applied to the data---it is not expected to communicate with relocation or transaction managers and incur indeterminate time penalties. Consequently, the emphasis in the design of stream--handlers should be to ensure they introduce a constant latency in transmission and reception. \subsection{Call--oriented Architecture} This architecture implicitly appreciates that the data being delivered is a call. Call--handlers usually add parameters to a remote procedure call. They do not form part of the message sent by the application object, but set its context. For example: \begin{itemize} \item Timestamps \- logging when messages are sent and received by adding an ``time--sent'' parameter to the remote procedure call. \item Accounting \- adding an account identifier to a remote procedure call so that the server can log charges for calls to a particular account. \item Transactions \- a transaction might consist of a number of calls to different servers, they could all be identified with a transaction identifier, to synchronize committing and aborting transactions as a whole. \item Authorization \- a Privilege Certificate could be attached to the call so that the server could check what rights and privileges the caller is allowed to exercise within the server's work--space. \end{itemize} A call--oriented architecture \textit{should} be implemented so that it has access to the parameters passed as part of the request. If this is the case, then as well as being able to add parameters, it would be possible to perform data transformations on parameter values that are part of the call. For example: \begin{enumerate} \item Representation conversions \begin{enumerate} \item Wholly The presentation objects implemented in remote procedure call systems change the data representation of the parameters of a call so that they can be transmitted over the network as a sequence of bytes, with the receiver being responsible for converting the byte--sequence to its local representation. It may be more efficient to convert to the receiver's format before the data is sent so that the server can use conversion methods available from its native operating system. \item Partly It might be the case that a server has a different data context for a particular data type: internationalization of text strings and currency formats could be converted prior to transmission. \end{enumerate} \item Pseudo--Objects Pseudo--objects are usually legacy systems that can be directly controlled by the client's remote procedure call infrastructure, but for the sake of uniformity, and to simplify re--engineering and relocation of services, they are provided with the same interface as remote servers. The operating system used by a distributed processing platform is itself a legacy system. An example of a pseudo--object that application programmers use is the database driver provided in the \emph{Java}\, DataBase Connectivity package, \mySrc{java.sql}. This pseudo--object implementation establishes and drives a connection with the database. Other examples of services that could be implemented as pseudo--objects are directory and naming services that are available through native operating systems. \item Stream A call--oriented architecture could also be used to transform data in the same way that a stream--oriented architecture could. Encryption, compression and checksum insertion could all be performed by marshalling the parameters using a data representation object to produce a sequence of bytes and then applying the stream operation to it. The output would be opaque and would replace the parameters. \end{enumerate} A call--oriented architecture is better suited to recovering from errors, since it is possible to determine which channel object is at fault and it can take measures to recover from the error. \subsection{Both architectures are needed} The stream--oriented architecture is ideal for delivering data at high speed with a determinate latency, the call--oriented architecture is ideal for communicating control information. This is a similar design problem that faced the developers of telephone networks and it was resolved, in the Integrated Services Digital Network, ISDN \cite{IEEE_1990}, by having a control channel manage the use of two bearer channels. Figure \ref{fig:odp-2sources} illustrates a request made to the server on a control channel and a reply being received on a broadband data channel. This sort of architecture could be used for controlling the delivery of ``pay--per--view'' television, where the RPC mechanism is used by an application to make a payment over a control network and the data is delivered on differently constructed channels over a broadband network possibly to different hardware. \begin{figure}[htbp] \begin{center} \includegraphics[angle=-90, keepaspectratio=1, totalheight=6 in]{odp-2sources.eps} \caption{Use of a RPC to connect to a broadband data source} \label{fig:odp-2sources} \end{center} \end{figure} Essentially the differences between the two architectures, with regard to the activation of the channel objects, are: \begin{itemize} \item synchronizing with the call, and \item the opacity of the call's contents \end{itemize} A stream--oriented architecture need not transmit data to the server in a contiguous block that represents the marshalled bytes of a message. A call--oriented architecture would have each channel object invoked with each call sent. A stream--oriented architecture only has access to the marshalled bytes that represent a message. A call--oriented structure sees the method that is invoked and the parameters for it. The most flexible architecture is the call--oriented one. But for implementing the transport objects, a stream--oriented architecture should be preferred. This means that the two architectures fall above and below the marshalling channel object, see figure \ref{fig:odp-call-stream}. \begin{figure}[htbp] \begin{center} \includegraphics[keepaspectratio=1, totalheight=4 in]{odp-call-stream.eps} \caption{The Marshalling Object: the boundary between call-- and stream--oriented architectures} \label{fig:odp-call-stream} \end{center} \end{figure} This figure attempts to place the channel objects in an Open Systems Interconnection model, \cite{OSI:intro}. Objects in the call--oriented architecture provide presentation layer services and, after the marshalling object, which reduces a call to a sequence of bytes, the session objects and finally the network transport object, a socket driver, can operate upon the data as a stream. The session layer objects would also perform stream--oriented encryption, but a presentation layer object would negotiate keys. Channel objects that are call--oriented will be called \emph{call--handlers} and those that are stream--oriented will be called \emph{stream--handlers}. The marshalling object is a call--handler and, if need be, it could invoked a number of time to render the data as a sequence of bytes. This would be useful for data security, since it may be necessary to encrypt the data and make it unintelligible to other call--handlers. \section{Stream--Oriented Architecture: Design} \label{sec:stream-oriented:design} \textit{Sun} with \emph{Java}\, and the socket factory technique have implemented what is described later as a simplex system, \myRef{sec:simplex}: each stream has two channel objects, one for sending---the output stream---and one for receiving---the input stream. There are two kinds of pairs: the client's, or, more precisely, the initiator's, pair and the server's, or acceptor's, pair. Although there is no explicit code to synchronize the two streams so that only one may be used at a time, it is not expected that an application can simultaneously send and receive. This limitation could be removed by a suitably designed channel object which could specify a different return address. \section{Call--Oriented Architecture: Design} As pointed out above, a message passed between a sender and receiver would negotiate four different channels: \begin{enumerate} \item Request \item Indication \item Response \item Confirmation \end{enumerate} but if the call is a cast of some kind, it need only have two: \begin{enumerate} \item Request \item Indication \end{enumerate} And some parts may not do anything, for example a request handler could log all calls made by the client, but the server need not record all the calls that are made upon it. When an application programmer makes use of a remote procedure call the infrastructure sends the message to the server and blocks the thread pending the arrival of the confirmation. One could also implement the channel objects in this way. This leads to two different architectures: \begin{itemize} \item Either: one channel object for each of request, indication, response and confirmation \- \emph{Simplex}. \item Or: two channel objects: one that requests and blocks pending the arrival of a confirmation; one that receives indications, invokes the service (and blocks waiting for it) and then sends the response \- \emph{Duplex}. \end{itemize} The latter will be discussed first. \subsection{Duplex: One Pair of Objects} One channel object for the initiator performs request and confirmation, another channel object performs indication and response for the acceptor. This means that every channel object has a bi--directional data--flow with the channel object below it, rather like the application object has with the distinct channels in figure \ref{fig:odp-4channels}. This has the attraction that should the initiator's channel objects for the request and the confirmation need to share state, then this is achieved implicitly because they are the same object. This has a problem with broadcasts, (or one--casts). The initiator's channel object (request and confirmation) needs to determine if it is to block or not and that would require this information be made available to the channel object, perhaps best supplied as a parameter to the method invocation. This information is required in \textit{CORBA}, the Interface Definition Language allows methods on interfaces to be marked as \mySrc{one way}, but is not part of the \emph{Java}\, specification; although it could be inferred if the definition of the remote method returns no result, in which case, the method can be invoked as a one--cast. However, if exceptions are returned by the remote server then the method is a call: the exception, whether raised or not, is a response. \subsection{Simplex: Two Pairs of Objects} \label{sec:simplex} If a channel object is implemented for each stage of the communication, then there are, at most, two pairs of objects: a request and confirmation pair at the initiator; an indication and response pair at the acceptor. Figure \ref{fig:odp-4channels} illustrates the engineering of this. Because the objects are distinct they will need to bind with one another should they need to share state: an example of the difficulties this might lead to can be seen when communicating and attempting to correct errors. The infrastructure would block the application programmer's thread of control pending the arrival of the confirmation message from the server containing the results of the call. Ordinarily, the confirmation channel objects will unravel the message returned by the server, instantiate the result for the application programmer's thread and unblock it. If there is an error in any of the channel objects, then it should be possible for the channel objects to attempt to clear the error and resend the message if need be. The problem then is that the channel objects in the request channel need to be synchronized with the error results received in the confirmation channel. If there is a failure in one of the channel objects in the indication channel, then they must be propagated via the response and confirmation channels. This is a difficult issue and is discussed later at greater length, \myRef{sec:exceptions}. The important difference between the duplex and simplex methods is the relationship to the state of the call. With the duplex model the request channel retains the state of the call awaiting an acknowledgement from the confirmation channel. What makes the duplex model retain state is that the channel objects invoke one another and form a stack of calls which can be associated with a thread. This can be emulated in the engineering of a simplex model without requiring the use of the stack, by passing a call identifier. This would only be of use if the channel objects were also to retain their internal state when they release control. \paragraph{Example: A Secured Message Delivery Service} As an example, an encryption and decryption service would require: \begin{enumerate} \item Request \- encrypt \item Indication \- decrypt \item Response \- encrypt \item Confirmation \- decrypt \end{enumerate} There are only two functions---encrypt and decrypt---but performed at four locations. It should be possible to provide just one pair of implementations---an \mySrc{Encryptor} and a \mySrc{Decryptor}---located differently. \begin{enumerate} \item \mySrc{Encryptor} \begin{itemize} \item Request channel object \- initiator \item Response channel object \- acceptor \end{itemize} \item \mySrc{Decryptor} \begin{itemize} \item Indication channel object \- acceptor \item Confirmation channel object \- initiator \end{itemize} \end{enumerate} \mySrc{Encryptor} and \mySrc{Decryptor} would both be implemented as stream--handler channel objects. Encryption is subject to replays of old messages unless the messages are time--stamped and sequenced. Usually, time--stamping and sequencing are implemented as part of the encryption and decryption objects, but with channel objects they can be provided separately. Time--stamping has two functions: \begin{enumerate} \item Request \- timestamp issue \item Indication \- timestamp check \item Response \- timestamp issue \item Confirmation \- timestamp check \end{enumerate} Two functions: \mySrc{StampIssuer}, \mySrc{StampChecker}; four locations: \begin{enumerate} \item \mySrc{StampIssuer} \begin{itemize} \item Request channel object \- initiator \item Response channel object \- acceptor \end{itemize} \item \mySrc{StampChecker} \begin{itemize} \item Indication channel object \- acceptor \item Confirmation channel object \- initiator \end{itemize} \end{enumerate} Similarly for a sequence number generator and checker. \mySrc{StampIssuer} and \mySrc{StampChecker} would both be implemented as call--handler channel objects. Because it is difficult to synchronize clocks in a distributed networks, some secure message delivery systems allow some skew on the clocks and use checksums to detect replayed messages. Only the acceptor's indication channel and the initiator's confirmation need deploy a \mySrc{ReplayDetector} object. It would be implemented as a call--handler generating a checksum from the marshalled data incoming as an indication or as a confirmation. \subsection{Simplex or Duplex} There is a greater similarity in the simplex architecture to the engineering underlying the messaging system than in the duplex architecture, but the duplex architecture has some attractive state--retention properties which should be emulated in a simplex architectureq. The rest of this discussion will concern itself with a simplex architecture that attempts to retain state across all four phases of a call. The other great attraction of the simplex architecture is that if the channel objects reside in different channels, then it is easier to relocate the channel. This would be especially useful when a call uses two different media as the example system in figure \ref{fig:odp-2sources} illustrated. \section{Service Invocation Semantics} From what has been said above, the \emph{Java}\, RPC mechanism has a stream--oriented architecture and what follows is a proposed call--oriented architecture for it. It should serve as an example of how channel objects could be deployed in other \textit{CORBA} RPC systems. The distributed processing infrastructure will have to determine the order in which the channel objects will be invoked. There are now two ways in which the methods of the channel objects can be invoked and how they should bind with one another. \begin{itemize} \item Either: have the request object invoke a method on the indication object and block waiting for the response: \emph{Peer--to--Peer} invocation. \item Or: have the request object perform its work and return a modified call object and return immediately: \emph{Service} invocation. \end{itemize} The peer--to--peer option requires the duplex architecture which has been dismissed, so only the service method of invocation is left. Peer--to--peer is very attractive, it would allow channel objects to be client--server pairs and thus be able to use the stub--compiler. Unfortunately, it would demand too much memory to stack all of the calls required to navigate complicated protocol stacks. The approach taken by \textit{Sun} in their own implementation of \mySrc{java.rmi.server.RemoteRef}, shipped as \mySrc{sun.rmi.server.UnicastRef} as part of the \emph{Java}\, Runtime Environment, is suitable for the simplex architecture proposed. Referring to the the code fragment given in the appendix \myRef{sec:stub}, the key method is \mySrc{java.rmi.server.RemoteRef.invoke()}. It is implemented along the lines given in the next code fragment: note the two comments indicating when the two types of channel objects should be invoked. \begin{verbatim} package sun.rmi.server; public class UnicastRef implements RemoteRef { public Object invoke(Remote r, reflect.Method m, Object[] p, long l) throws Exception { // *Request channel objects should be invoked now* // Establish a connection using information in Remote r // Get the streams associated with the connection // Marshall data onto the stream // Execute the call // *Confirmation channel objects should be invoked now* // Unmarshall // Release the connection // Return the result } } \end{verbatim} \subsection{Wrappers: Request and Response Channels} The channel objects would be invoked serially and would be passed the same parameters as \mySrc{invoke()}, collectively call these a \mySrc{Message} object. The request channel objects would return a \mySrc{Message} object, but these would usually have the original message as one of its parameters. This is a simple encapsulation procedure and is illustrated in figure \ref{fig:odp-channel-message-1}, where a time--stamping channel object has been passed the application object's message. The application object wants to invoke method \mySrc{query()}, the channel object passes this message as a parameter of its own message. That message invokes method \mySrc{stampedAt()}. The two message objects might duplicate target and return addresses, it should be possible to remove this redundancy, but it should still be possible to specify a different target address and a different return address if need be. \begin{figure}[htbp] \begin{center} \includegraphics{odp-channel-message-1.eps} \caption{A message containing another message} \label{fig:odp-channel-message-1} \end{center} \end{figure} \subsection{Unwrappers: Indication and Confirmation Channels} The objects located in the indication and confirmation channels would need a skeleton rather like that described above, \myRef{sec:inflexible}, either a dispatcher or use a reflective invocation on the channel object's service implementation. After the service has completed processing, it would return its results to the skeleton which would then return the message object to the infrastructure. \subsection{Counterparts and Associates} Some channel objects will have counterparts in the remote channel, for example, a time--stamping request channel object should have a time--stamp checking indication channel object, but some may not, for example a request logging channel object placed in a server's indication channel. If a channel object sends then its counterpart receives and \textit{vice--versa}. A channel object can have an associate in a local channel. A request channel object could have an associate in the confirmation channel. If a channel objects sends then its associate receives and \textit{vice--versa}. Figure \ref{fig:odp-counterparts} should help to clarify this. \begin{figure}[htbp] \begin{center} \includegraphics[angle=-90, keepaspectratio=1, totalheight=6 in]{odp-counterparts.eps} \caption{Counterpart and Associate Channel objects} \label{fig:odp-counterparts} \end{center} \end{figure} \begin{enumerate} \item Marshalling and Unmarshalling Looking at the marshalling and unmarshalling objects: the marshalling object in the client's request channel has a counterpart in the server's indication channel and an associate in the client's confirmation channel. \item Timing The top layer of channel objects are used for timing: time--stamping and time--checking, they comprise a full complement, where each channel object has an associate and a complement. \item Usage The second layer of channel objects is used to record usage statistics and only logs requests and responses, so they have no associates but a counterpart. \end{enumerate} Clearly, this might prove to be cumbersome, it might be simpler to insist that all channel objects have a counterpart and an associate and have a default implementation which just copies the message over. \subsection{Exception Handling} \label{sec:exceptions} Any of the channel objects can raise an exception, but part of the function of channel objects is to attempt to clear exceptions, for example: \begin{itemize} \item relocation objects would obtain new addresses for servers, \item key management objects would obtain new keys in the event of expiry. \item authorization managers could obtain new privileges. \end{itemize} The channel objects have to retain some state that would allow them to act upon exceptions. This would require that they place state in a common object that associated channel objects can access. The state would need to be stored with an identifier unique to the call. \begin{enumerate} \item Exception Handling in One Channel \begin{enumerate} \item Clear When one channel object raises an exception, the infrastructure signals the other channel objects in the same channel, in the reverse order in which they were invoked, asking them to attempt to clear the exception. \item Uncleared: Undo If the exception cannot be cleared then the channel objects should be signalled to undo their previous actions. \item Cleared: Undo and Redo If the exception can be cleared then the channel objects may need to undo their previous actions and be allowed to redo them. \end{enumerate} Redo is a distinct operation, because it may allow the implementation to be optimized for error recovery. Co--ordinating this is a little difficult. There are two sequences: \begin{enumerate} \item Attempt to clear then undo and then redo Each object in turn in ascending order (\textit{i.e.}\, reverse order to invocation) attempts to clear the exception, if any one succeeds then the undo operation is invoked in ascending order to the top of the channel. Then the redo operation is invoked in descending order. \item Attempt to clear and undo and then redo Each object in turn in ascending order (\textit{i.e.}\, reverse order to invocation) attempts to clear the exception and performs an undo. If any one succeeds then the undo is invoked on the remaining objects in ascending order and then the redo operation is invoked in descending order. \end{enumerate} The former might prove more efficient if errors are expected to be cleared, the latter if not. The idea is communicated in figure \ref{fig:odp-exceptions}, but the details of invocation are not. \begin{figure}[htbp] \begin{center} \includegraphics{odp-exceptions.eps} \caption{Exception handling in one channel} \label{fig:odp-exceptions} \end{center} \end{figure} If the latter scheme (Attempt to Clear and Undo Simultaneously) is used: channel object $C$ raises an exception, it undoes its action, passes back the original message it received to object $B$ which also undoes its action and returns the original message it received to $A$. $A$ clears the exception and redoes its action and passes the message on to $B$ which also redoes its action and thence to $C$. \item Exception Handling Across Channels If a receiving counterpart raises an exception, it should be signalled to the sender: this would be achieved by the receiving counterpart sending a message to its sending associate to raise an exception with its receiving counterpart. \emph{Java}\, already has a proven mechanism for this, objects of class \mySrc{Exception} can be contained in objects of class \mySrc{RemoteException}. \begin{enumerate} \item Exception raised in the Indication Channel As an example, see figure \ref{fig:odp-exceptions-1}, a request channel object sends a message which raises an exception in the indication channel, the channel object in the indication channel raises an alert with its associate in the response channel, which sends an exception to its counterpart in the confirmation channel. \begin{figure}[htbp] \begin{center} \includegraphics[angle=-90, keepaspectratio=1, totalheight=6 in]{odp-exceptions-1.eps} \caption{Exception handling across channels} \label{fig:odp-exceptions-1} \end{center} \end{figure} Points to note in figure \ref{fig:odp-exceptions-1} are: \begin{enumerate} \item The indication channel is cleared down \item The server receives no message \item The response channel propagates the exception \item The objects in the confirmation channel can signal their associates in the request channel. \end{enumerate} It might be possible for the request channel objects to invoke the same procedure as portrayed in figure \ref{fig:odp-exceptions}, clear the exception and re--send the message, this would require that they have access to the message as issued by the client. \item Exception raised in the Confirmation Channel If a response channel object sends a message which raises an exception in the confirmation channel, then the confirmation channel object would signal its associate in the request channel, which may, if it has retained state, \textit{i.e.}\, the message it sent, be able to clear the exception and re--send without intervention by the client application object. \end{enumerate} \end{enumerate} \subsection{Interfaces and Classes} \subsubsection{A \mySrc{Message} Class} A simple message class is needed to contain the parameters passed to the \mySrc{invoke()} method and to all the channel objects. \begin{verbatim} package java.lang.rmi.channel; import java.rmi.*; import java.lang.reflect.*; public class Message { Remote remote; Method method; Object[] parms; long hash; public Message(Remote r, Method m, Object[] p, long h) { remote = r; method = m; parms = p; hash = h; } } \end{verbatim} \subsubsection{Channel Objects} \paragraph{Basic Interface} Channel objects then all have the same interface and it is their location which determines their function, \textit{i.e.}\, whether they wrap or unwrap. \begin{verbatim} package java.lang.rmi.channel; public interface Handler { public Message clear(Message m, Exception e) throws Exception; public Message todo(Message m) throws Exception; public Message undo(Message m, Exception e) throws ClearedException; public Message redo(Message m) throws Exception; } \end{verbatim} Both methods of clearing exceptions are covered in this interface because there is a separate \mySrc{clear()} method. \paragraph{Exceptions} The \mySrc{undo()} method throws an exception to indicate it has cleared the exception it was passed. \begin{verbatim} package java.lang.rmi.channel; public class ClearedException extends Exception { public ClearedException(String s) { super(s); } public ClearedException(String s, Exception ex) { super(s, ex); } } \end{verbatim} Other exceptions which might make processing more decisive: \begin{enumerate} \item Unclearable It might also prove useful to have \mySrc{clear()} raise an exception that the exception cannot be cleared and this would allow the infrastructure to request a re--send decision from the application object. \item Rebind It might also prove useful if the channel objects can demand a rebind and force the destruction of their channel. This would be useful for a relocation channel object. \end{enumerate} \paragraph{Handler Specification} The next is an abstract class that provides a container to hold the channel objects which would be loaded by the infrastructue. It might be wise, for security reasons, to implement the class more fully and make the \mySrc{getHandler()} method final. The class also provides a means of obtaining associates and counterparts. \begin{verbatim} package java.lang.rmi.channel; import java.rmi.*; public abstract class Handlers { public static final int REQUEST = 1; public static final int INDICATION = 2; public static final int RESPONSE = 3; public static final int CONFIRMATION = 4; public Handlers() { ; } Handler getHandler(int identity) { return null; } Remote getCounterPart(int identity) { return null; } Object getAssociate(int identity) { return null; } } \end{verbatim} \paragraph{Channels} Channels themselves would be simple containers probably implemented with a vector object which can be easily iterated in the \mySrc{invoke()} operation. \begin{verbatim} package java.lang.rmi.channel; import java.rmi.*; public abstract class Channel { private int identity = -1; private Handlers handlers; public Channel(int id, Channels h) { identity = id; handlers = h; } } \end{verbatim} \subsubsection{Channel Objects: Binding and Invoking Methods} \paragraph{Binding} Channel objects would need to bind with one another, so that they can exchange parameters. The channel object methods are defined on one interface, they can therefore suppport other methods on another service interface. \paragraph{Invocation} When a channel object is passed a message object, it will need to invoke a method on its counterpart. The easiest way for it to do this is to use the same stub and skeleton mechanism for dispatching a call as remote procedure calls use. As can be seen in figure \ref{fig:odp-channel-message-1}, where one channel object adds enough parameters to the message so that when it is received by its counterpart, it can use a local \mySrc{invoke()} method to pass the parameters and obtain the resulting message object from a local service interface. If channel objects have to invoke methods upon one another they can either send them with the message or they can go ``out--of--band'' and communicate them directly if they support the \mySrc{java.rmi.Remote} interface using their own channel. \section{Summary} \subsection{Previous Work} \emph{Java}\, certainly has the functionality to implement channel objects. The author has already implemented a similar scheme (using a duplex, peer--to--peer architecture) for a variant of the \textit{ANSAware}, \cite{web:ansa}, known as \textit{DAIS}, the distributed processing platform produced by \textit{ICL}, \cite{web:icl}, which was \textit{CORBA} compliant. That implementation, written wholly in \textit{C}, \cite{Kernighan:CPL78}, was able to support a \textit{Kerberos}, \cite{sec:kerberos}, authentication and confidentiality service invisibly to the application programmer, the channels were constructed according to a template specified by an environment variable. The author has developed a prototype application which can support the Transport Level Security protocol. (A fairly rigorous treatment of both it and the \emph{Kerberos} protocol is given in \cite{tls:eaves}.) A simple channel object is put in place that negotiates the keys, whilst a custom socket is used to perform the encryption. \textit{ANSAWare} evolved to a product known as \textit{Reflective Java}, \cite{sft:reflect:java}, which supported channel objects, but was designed to provide support to application objects---presentation of data and so forth---and not to provide system services. \subsection{Difficulties of Binding} The principal difficulty faced in these prototypes is that there is no simple or well--defined mechanism for expressing how two parties should bind with one another. \textit{ANSAware} had a well--developed binding model, \cite{sft:ansaware}, but its final implementation had a limited number of quality of service parameters. \emph{Java}\, now has a very well--developed security architecture, \cite{sec:java:security}, but currently seems to have no means of specifying the degree of security \emph{required}, it seems to be principally oriented towards setting permissions. (Of course, a permission one does not have is a requirement.) The \mySrc{SecurityManager} would appear to be the best--placed component of the \emph{Java}\, security architecture to negotiate and configure channels. It will be difficult for any distributed processing platform to achieve any degreee of acceptance in an enterprise--wide data processing environment if it is not possible to implement many of the logging services that have long been available in mainframe systems. The most pertinent example of which is \textit{CICS} \cite{Wipfler87}, the Customer Information Control System, which was originally a messaging system, but was extended to become a transaction monitor that could manage database enquiries. Hopefully, with channel objects in place, remote procedure calls could do the same across the World--Wide Web.
\section{introduction} The cosmic censorship hypothesis (CCH) is one of the most important open problems in classical gravity (Penrose 1979). The CCH roughly says that the physically reasonable space-time contains no naked singularity. Since the CCH asserts the future predictability of the space-time, it is so helpful that several theorems on black holes have been proved under the assumption of CCH (Hawking and Ellis 1973). In spite of all the effort, the censorship has not yet been proved. In fact, there have been discovered several solutions which have naked singularities with matter content that satisfies energy conditions. Then the curvature strength of singularities was defined in a hope that weak convergence would reveal the extendibility of the space-time in a distributional sense. In this context, Tipler (1977) defined the strong curvature condition (SCC), while Kr\'olak (1987) defined weaker condition, which we call the limiting focusing condition (LFC). One of the most important examples which have naked singularities is the Lema\^{\i}tre-Tolman-Bondi (LTB) solution. This solution describes the spherical collapse of an inhomogeneous dust ball. It has been proved that this solution has naked singularities from generic initial data. The naked singularities in this solution are classified to ``shell-crossing'' (Yodzis, Seifert and M\"uller zum Hagen 1973) and ``shell-focusing'' (Eardley and Smarr 1979, Christodoulou 1984) singularities. Newman (1986) showed that, for a null geodesic from the shell-crossing singularity, neither SCC nor LFC is satisfied. It is widely believed that the shell-crossing singularities would be harmless because they would be dealt with in some distributional sense. Newman (1986) also showed that, for a null geodesic from the shell-focusing singularity which results from generic smooth initial data, not SCC but LFC is satisfied. Hence, shell-focusing singularities will be more serious to CCH than shell-crossing singularities. It is important that the strength of the singularity is determined by the curvature divergence not only on the null geodesic but also on the timelike geodesic. Recently, Deshingkar, Joshi and Dwivedi (1999) showed that both SCC and LFC are satisfied for the timelike geodesics which terminate at the shell-focusing singularity. It might be thought that, since the dust is a pressureless fluid, there appears naked singularity which satisfies LFC. As an extension of the LTB solution, we will consider a spherically symmetric space-time with a fluid that has only tangential pressure. Magli (1997, 1998) solved an explicit solution with the mass-area coordinates. Here we give a formalism to examine the existence of naked central singularity and its curvature strength. Next we proceed further by defining the ``gravity dominance condition''. After that we discuss several examples. We follow the sign conventions of the textbook by Misner, Thorne and Wheeler (1973) about the metric, Riemann and Einstein tensors. We use the units with $c=G=1$. \section{metric functions and occurrence of naked singularity} In a spherically symmetric space-time, the line element is written in the diagonal form as \begin{equation} ds^2=-e^{2\nu(t,r)}dt^2+e^{2\lambda(t,r)}dr^2+R^2(t,r)(d\theta^2 +\sin^2\theta d\phi^2). \label{eq:lineelement} \end{equation} Using the comoving coordinates, the stress-energy tensor $T^{\mu}_{~\nu}$ with vanishing radial pressure is of the following form: \begin{equation} T^{\mu}_{~\nu}= \pmatrix{ -\epsilon & 0 & 0 & 0 \cr 0 & 0 & 0 & 0 \cr 0 & 0 & \Pi & 0 \cr 0 & 0 & 0 & \Pi \cr }, \end{equation} where $\epsilon(t,r)$, $\Pi(t,r)$ are the energy density and the tangential pressure, respectively. From the Einstein equation and the equation of motion for the matter, we obtain \begin{eqnarray} m&=&F, \label{eq:mconserve}\\ \epsilon&=&\frac{F^{\prime}}{4\pi R^2 R^{\prime}}, \label{eq:energydensitycomoving}\\ e^{2\lambda}&=&R^{\prime 2}h^2, \label{eq:grr} \\ \nu^{\prime}&=&-\frac{1}{h}h_{,R}R^{\prime}, \label{eq:lapse}\\ \dot{R}^2 e^{-2\nu}&=&-1+\frac{2F}{R}+\frac{1}{h^2}, \label{eq:energyofparticle} \end{eqnarray} where an arbitrary function $F=F(r)$ is the conserved Misner-Sharp mass (Misner and Sharp 1973). The dot and prime denote the partial derivatives with respect to $t$ and $r$, respectively. We have introduced a function $h=h(r,R)\ge 0$ as \begin{equation} \Pi=-\frac{R}{2h}h_{,R}\epsilon, \label{eq:eos} \end{equation} where the comma denotes the partial derivative. We should note that the definition of $h$ is slightly different from Magli (1997, 1998)'s notation. The dust limit is given by $h=h(r)$. Eqs.~(\ref{eq:lapse}) and (\ref{eq:energyofparticle}) are coupled and cannot be integrated explicitly. We can express the metric functions explicitly by introducing the mass-area coordinate system \begin{equation} ds^2 = -A(m,R) dm^2- 2B(m,R) dmdR -C(m,R) dR^2 +R^2 (d\theta^2 +\sin^2\theta d\phi^2), \end{equation} which was introduced by Ori (1990). Since the derivation of the explicit solution was described in Magli (1998), here we only present the results: \begin{eqnarray} A&=&H\left(1-\frac{2m}{R}\right), \\ B&=&-\frac{\sqrt{H}}{h|u|}, \\ C&=&\frac{1}{u^2}, \end{eqnarray} where \begin{eqnarray} \sqrt{H}&=&\frac{R^{0}_{,m}h(m,R^{0}(m))}{|u^{0}|} +\int^{R}_{R^{0}}\frac{h}{x}\left(1+\frac{x}{2}\left(\frac{1} {h^2}\right)_{,m}\right)\left(-1+\frac{2m}{x}+\frac{1}{h^2} \right)^{-3/2}dx, \label{eq:rootH}\\ u&\equiv&\frac{dR}{d\tau}=\pm\sqrt{-1+\frac{2m}{R}+\frac{1}{h^2}}, \label{eq:velocityofshell}\\ u^{0}(m)&\equiv&\pm\sqrt{-1+\frac{2m}{R^{0}(m)} +\frac{1}{h^2(m,R^{0}(m))}}, \\ \label{eq:energydensityma} R^{0}(m)&\equiv& R(0,F^{-1}(m)), \end{eqnarray} the energy density is given as \begin{equation} \epsilon=\frac{h}{4\pi R^2|u|\sqrt{H}}, \end{equation} and we have assumed $\epsilon\ge 0$. The shell-crossing singularity is the one characterized by $R^{\prime}=0$ and $R>0$, while the shell-focusing singularity is the one characterized by $R=0$. Newman (1986) showed that the shell-crossing singularities do not satisfy even LFC for a null geodesic. Christodoulou (1984) showed that noncentral ($r>0$ or $m>0$) shell-focusing singularities are not naked. Therefore we concentrate on central ($r=0$ or $m=0$) shell-focusing singularities. If and only if the singularity is naked, there exists an outgoing null geodesic which emanates from the singularity. In the mass-area coordinates, we can derive the root equation which probes the existence of such a geodesic as follows. The radial null rays are determined by the equation \begin{equation} \frac{dR}{dm}=\frac{-B\mp\sqrt{H}}{C} =\sqrt{H}|u|\left(\frac{1}{h}\mp|u|\right). \label{eq:drdm} \end{equation} We should note that the upper sign refers to an outgoing null ray in a collapsing phase and an ingoing null ray in an expanding phase {\em at the same time}. Similarly, the lower sign refers to an ingoing null ray in a collapsing phase and an outgoing null ray in an expanding phase {\em at the same time}. Hereafter we mainly concentrate on a collapsing phase. Here we define \begin{equation} y\equiv\frac{R}{2m^{\beta}}, \end{equation} where $\beta$ is determined by requiring that $y$ has a positive finite limit $y_0$ along the null geodesic. Then, the regular center corresponds to $\beta\le 1/3$. If we assume that the energy density at the center is positive, the regular center corresponds to $\beta= 1/3$. Note that we will only consider naked singularities with such $\beta>1/3$. It is noted that we will assume the existence of every limit through this paper in a sense including $\pm \infty$. Then, from the l'Hospital's rule, \begin{equation} y_{0}=\lim_{m\to0}\frac{R}{2m^{\beta}} =\lim_{m\to0}\frac{m^{1-\beta}}{2\beta}\frac{dR}{dm} =\left.\lim_{m\to0}\frac{m^{1-\beta}}{2\beta}\sqrt{H}|u| \left(\frac{1}{h}\mp |u|\right)\right|_{R=2y_0m^{\beta}}. \label{eq:rooteqma} \end{equation} Therefore, we obtain the root equation for the existence of the null geodesic from the central singularity \begin{equation} y_0 =\frac{1}{2\beta}\lim_{m\to0} \left[m^{3(1-\beta)/2}\sqrt{H} \sqrt{\left(-1+\frac{1}{h^2}\right)m^{-(1-\beta)}+\frac{1}{y_0}} \left(\frac{1}{h}\mp \sqrt{\frac{m^{1-\beta}}{y_0}-1+\frac{1}{h^2}} \right)\right], \label{eq:rooteq} \end{equation} where the limit is taken along $R=2y_{0}m^{\beta}$. This equation was first derived by Magli (1998). As seen in this equation, the existence of a future-directed outgoing null ray from the singularity in a collapsing phase requires \begin{equation} \frac{1}{3}< \beta\le 1. \end{equation} From $u^2\ge0 $ and $0<y_{0}<\infty$, \begin{equation} \lim_{m\to 0}h\le \left\{\begin{array}{ll} 1&\mbox{for $\beta<1$} \\ (1-y^{-1}_{0})^{-1/2}&\mbox{for $\beta=1$} \end{array}\right., \end{equation} where the limit is taken along the null ray which emanates from the singularity. Here we should note that, if the singularity is {\em critically naked}, i.e., if \begin{equation} \lim_{m\to 0}\frac{2m}{R}=1 \end{equation} holds along the null ray, higher order analysis is needed. \section{curvature condition along null geodesic} We consider a radial null geodesic which emanates from or terminates at the naked singularity. We prepare a parallely propagated tetrad $E_{i}:(i=1,2,3,4)$ with $E_{1}\cdot E_{1}=E_{2}\cdot E_{2} =-E_{3}\cdot E_{4}=-E_{4}\cdot E_{3}=1$, all other products vanish and $E_{4}$ is equal to the tangent vector $k^{\mu}$ of the null geodesic. In a spherically symmetric space-time, for the tetrad components of the Weyl tensor, \begin{equation} C^{m}_{~4n4}=0 \end{equation} holds for $m,n=1,2$. Define \begin{equation} p\equiv\lim_{\lambda\to +0}\lambda^{\alpha} R_{44}, \label{eq:defp} \end{equation} where $R_{44}$ is defined by \begin{equation} R_{44}\equiv R_{\mu\nu}k^{\mu}k^{\nu} \label{eq:defR44} \end{equation} and $\lambda$ is the affine parameter such that $\lambda\to +0$ corresponds to an approach to the singularity. Then, from Clarke and Kr\'olak (1985) and Clarke (1993), \begin{lm} For the radial null geodesic which emanates from or terminates at the singularity in the spherically symmetric space-time: SCC is satisfied if $p$ is positive for $\alpha=2$, and not satisfied if $p$ is equal to $0$ for $\alpha<2$; LFC is satisfied if $p$ is positive for $\alpha=1$, and not satisfied if $p$ is equal to $0$ for $\alpha<1$. \end{lm} Since the null geodesic is given as \begin{equation} k^{R}=\frac{-B \mp \sqrt{H}}{C}k^{m}, \label{eq:nullcondition} \end{equation} we obtain, from the form of the stress-energy tensor, \begin{equation} R_{44}=8\pi \epsilon u^2 H (k^{m})^{2}. \end{equation} Then, \begin{equation} \lambda^2 R_{44} = \frac{1}{2} |u|h\sqrt{H} \left(\frac{2m}{R}\right)^2 \left(\frac{\lambda}{m}\frac{d m}{d\lambda} \right)^2 \approx\frac{\beta q^2}{y_{0}} m^{1-\beta}\frac{h^2}{1\mp h|u|} \end{equation} holds, where $q$ is defined as \begin{equation} q\equiv \lim_{m\to 0} \frac{d\ln m}{d\ln\lambda} \end{equation} and we have used Eq.~(\ref{eq:rooteqma}). We should note that, for $(\beta,y_{0})\ne (1,1)$, \begin{equation} 0<\lim_{m\to 0}\frac{h^2}{1-h|u|}<\infty \end{equation} and \begin{equation} 0\le\lim_{m\to 0}\frac{h^2}{1+h|u|}<\infty \end{equation} hold, where the equality holds only when \begin{equation} \lim_{m\to 0}h=0 \end{equation} is satisfied. Therefore, \begin{equation} R_{44}\propto \lambda^{q(1-\beta)-2} \label{eq:R44} \end{equation} holds for the outgoing null geodesic with $0<q<\infty$ and $ (\beta,y_{0})\ne (1,1)$, and for the ingoing null geodesic with $0<q<\infty$, $ (\beta,y_{0})\ne(1,1)$ and $\lim_{m\to 0}h\ne 0$. In summary, we present the following theorems: \begin{th} For the outgoing radial null geodesic which emanates from the noncritically naked singularity with $0<q<\infty$: if and only if $1/3<\beta<1$ and $(1-\beta)^{-1}<q<\infty$ are satisfied, neither SCC nor LFC holds, if and only if $1/3<\beta<1$ and $0<q\le (1-\beta)^{-1}$ are satisfied, not SCC but only LFC holds, and if and only if $\beta=1$ is satisfied, both SCC and LFC hold. \end{th} \begin{th} For the ingoing radial null geodesic which terminates at the noncritically naked singularity with $0<q<\infty$ and $lim_{m\to 0}h\ne 0$: if and only if $1/3<\beta<1$ and $(1-\beta)^{-1}<q<\infty$ are satisfied, neither SCC nor LFC holds, if $1/3<\beta<1$ and $0<q\le (1-\beta)^{-1}$ are satisfied, not SCC but only LFC holds, and if and only if $\beta=1$ is satisfied, both SCC and LFC hold. \end{th} In order to estimate $q$, we must solve the null geodesic equation \begin{equation} \frac{d}{d\lambda}\left(\sqrt{H}k^{m}\right) \pm \frac{1}{2}\left[A_{,R}+2 B_{,R} |u| \sqrt{H}\left(\frac{1}{h}\mp |u|\right) +C_{,R} u^2 H \left(\frac{1}{h}\mp |u|\right)^2 \right](k^{m})^{2}=0, \label{eq:nullgeodesiceqma} \end{equation} where we have used the null condition (\ref{eq:nullcondition}). From Eq.~(\ref{eq:rootH}), we obtain \begin{equation} \sqrt{H}_{,R}=\frac{h}{2|u|^3}\left(\frac{2}{R} +\left(\frac{1}{h^2}\right)_{,m}\right). \end{equation} Using this, the following expressions are derived. \begin{eqnarray} A_{,R}&=&\frac{2m}{R^2}H+\frac{\sqrt{H}}{|u|^3}h \left(1-\frac{2m}{R}\right) \left(\frac{2}{R}+\left(\frac{1}{h^2} \right)_{,m}\right), \\ B_{,R}&=&-\frac{1}{2|u|^4}\left(\frac{2}{R}+\left(\frac{1}{h^2} \right)_{,m}\right) -\frac{\sqrt{H}}{|u|^3}\frac{1}{h} \left[\left(1-\frac{2m}{R}\right)\frac{h_{,R}}{h}+ \frac{m}{R^2}\right],\\ C_{,R}&=&\frac{2}{u^4}\left(\frac{m}{R^2}+\frac{1}{h^2} \frac{h_{,R}}{h}\right). \end{eqnarray} Using these expressions, we finally obtain the radial null geodesic equation for $m\to 0$ in the explicit form \begin{equation} \frac{d^2m}{d\lambda^2}=\left[1-\beta -\frac{1}{2}\frac{h^2}{1\mp h|u|} \left(\frac{2m}{R}+\frac{d}{d\ln m}\frac{1}{h^2}\right)\right] \frac{1}{m}\left(\frac{dm}{d\lambda}\right)^2, \label{eq:nullgeo} \end{equation} where the ordinary derivative is taken along $R=2y_{0}m^{\beta}$. In evaluating the right hand side of Eq.~(\ref{eq:nullgeo}), we have used \begin{eqnarray} R&\approx&2y_{0}m^{\beta}, \label{eq:R}\\ |u|&\approx&m^{(1-\beta)/2}\sqrt{\frac{1}{y_{0}} +m^{-(1-\beta)}\left(\frac{1}{h^2}-1\right)}, \label{eq:u}\\ \sqrt{H}&\approx&\beta\frac{R}{m}\frac{h}{|u|(1\mp h|u|)}. \label{eq:rtH} \end{eqnarray} On the other hand, if $m$ is proportional to $\lambda^{q}$ along the null geodesic, the following equation holds: \begin{equation} \frac{d^2m}{d\lambda^2}=\left(1-\frac{1}{q}\right)\frac{1}{m} \left(\frac{dm}{d\lambda}\right)^2. \label{eq:mlambdaq} \end{equation} Comparing Eqs.~(\ref{eq:nullgeo}) and (\ref{eq:mlambdaq}), we can determine $q$. Therefore the curvature divergence along the null geodesic is determined only from $\beta$ and $h$ along the geodesic. \section{gravity dominance condition} \label{sec:gravitydom} Here we define the gravity dominance condition (GDC) and the gravity-dominated singularity as follows: \newtheorem{df}{Definition} \begin{df}[Gravity Dominance Condition] For the geodesic which emanates from or terminates at the singularity, we have \begin{equation} \lim_{m\to 0}\frac{R}{2m}\left(\frac{1}{h^2}-1\right)= 0. \label{eq:condition1} \end{equation} \end{df} \begin{df}[Gravity-Dominated Singularity] A singularity is said to be gravity-dominated if and only if GDC is satisfied for every causal geodesic which emanates from or terminates at the singularity. \end{df} GDC is satisfied for the geodesic which emanates from or terminates at a very wide class of naked singularities. Furthermore, if the gravitational collapse of physical matter from regular initial data results in the central naked singularity formation, GDC is satisfied for the null geodesic, at least within our knowledge. The important example is the central singularity in the collapse of the spherical cluster of counterrotating particles which will be discussed in Sec. VI. If GDC is satisfied for the null geodesic, the collapse is induced dominantly by the gravitational potential (see Eq.~(\ref{eq:energyofparticle}) or (\ref{eq:velocityofshell})) and the null geodesic equation is controlled only by the gravitational potential (see Eq.~(\ref{eq:nullgeo}). The latter can be shown by the following proposition. \begin{pr} If GDC is satisfied, then \begin{equation} \lim_{m\to 0}\frac{R}{2m}\frac{d}{d\ln m}\frac{1}{h^2}=0 \label{eq:condition2} \end{equation} holds. \end{pr} {\em Proof.} We use $R\approx 2y_{0}m^{\beta}$ along the null geodesic. Then, for $\beta<1$, the l'Hospital's rule applies because we have assumed the exsitence of the limit. From condition (\ref{eq:condition1}), the proposition holds. For $\beta=1$, we set $f \equiv h^{-2}-1$. Then condition (\ref{eq:condition1}) implies $\lim_{x\to 0}f(x)=0$. From the mean value theorem, there exists $c\in (0,x)$ for any $x>0$ such that \[ \left|c f^{\prime}(c)\right| =\left|c\frac{f(x)}{x}\right|\le|f(x)|. \] Because we have assumed the existence of the limit, it must be zero.\hfill $\Box$ If GDC is satisfied for the null geodesic, Eqs.~(\ref{eq:u}) and (\ref{eq:rtH}) become \begin{equation} |u|\approx y_{0}^{-1/2}m^{(1-\beta)/2}, \end{equation} and \begin{equation} \sqrt{H}\approx\left\{ \begin{array}{ll} 2\beta y_{0}^{3/2}m^{-3(1-\beta)/2},& \mbox{for $\beta<1$}\\ \displaystyle\frac{2 y_{0}^{2}}{y_{0}^{1/2}\mp 1}, & \mbox{for $\beta=1$ and $y_{0}\ne1$}\\ \end{array}\right.. \end{equation} For $\beta<1$, since Eq.~(\ref{eq:nullgeo}) becomes \begin{equation} \frac{d^2m}{d\lambda^2} =(1-\beta)\frac{1}{m}\left(\frac{dm}{d\lambda}\right)^2, \end{equation} we obtain \begin{equation} q=\frac{1}{\beta}. \end{equation} Then, $1<q<3$ and $R\propto \lambda$ hold. Eq.~(\ref{eq:R44}) becomes \begin{equation} R_{44}\propto \lambda^{-3+\beta^{-1}}. \end{equation} Therefore SCC is not satisfied. LFC is satisfied for $1/2\le \beta <1$, while LFC is not satisfied for $1/3<\beta<1/2$. For $\beta=1$ and $y_{0}\ne 1$, Eq.~(\ref{eq:nullgeo}) becomes \begin{equation} \frac{d^2m}{d\lambda^2} =-\frac{1}{2(y_{0}^{1/2}\mp 1)y_{0}^{1/2}} \frac{1}{m}\left(\frac{dm}{d\lambda}\right)^2. \end{equation} Therefore we obtain \begin{equation} q=\frac{2(y_{0}^{1/2}\mp 1)y_{0}^{1/2}} {2(y_{0}^{1/2}\mp 1)y_{0}^{1/2}+1}, \end{equation} Then, $0<q<1$ and $R\propto \lambda^{q}$ hold. Eq.~(\ref{eq:R44}) becomes \begin{equation} R_{44}\propto \lambda^{-2}. \end{equation} Therefore both SCC and LFC are satisfied. In summary, we present the following theorems: \begin{th} Suppose that GDC is satisfied for a radil null geodesic which emanates from or terminates at the noncritically naked singularity. If and only if $1/3<\beta<1/2$ is satisfied, neither SCC nor LFC holds, if and only if $1/2\le \beta<1$ is satisfied, not SCC but only LFC holds, and if and only if $\beta=1$ is satisfied, then both SCC and LFC hold, for the radial null geodesic which emanates from or terminates at the singularity. \end{th} \begin{th} Suppose that GDC is satisfied for a radil null geodesic which emanates from or terminates at the noncritically naked singularity. Along the radial null geodesic, \[ \lim_{m\to 0}\frac{R}{\lambda} \] is nonzero finite value or positive infinity. If and only if the limit converges, SCC does not hold, and if and only if the limit diverges, both SCC and LFC hold, for the null geodesic. \end{th} \section{curvature condition along timelike geodesic} Here we consider a timelike geodesic which terminates at the singularity. We prepare a parallely propagated tetrad $E_{i}$ with $E_{1}\cdot E_{1}=E_{2}\cdot E_{2}=E_{3}\cdot E_{3}=-E_{4}\cdot E_{4}=1$, all other products vanish and $E_{4}$ is equal to the tangent vector $k^{\mu}$ of the timelike geodesic. We can define $p$ and $R_{44}$ by Eqs.~(\ref{eq:defp}) and (\ref{eq:defR44}), respectively, also for the timelike geodesic. From Clarke and Kr\'olak (1985) and Clarke (1993), the following lemma holds. \begin{lm} For the timelike and null geodesic which emanate from or terminate at the singularity: SCC is satisfied if $p$ is positive for $\alpha=2$; LFC is satisfied if $p$ is positive for $\alpha=1$. \end{lm} It seems to be cumbersome to examine the curvauture diveregnce along all possible timelike geodesics. Then, we consider the simplest timelike geodesic, i.e., $r=0$. It is easy to find that $r=0$ is a timelike geodesic when the center is regular. As a matter of convenience, we adopt the coordinate system (\ref{eq:lineelement}). Along $r=0$, the $R_{44}$ is calculated as \begin{equation} R_{44}=4\pi \epsilon = \frac{F^{\prime}}{R^2R^{\prime}}, \label{eq:R44timelike} \end{equation} where we have used $\Pi=0$ at the regular center which will be seen in Sec. VII. We will consider the situation in which the central singularity occurs at $t=0$ from the regular initial data at $t=t_{0}<0$. We choose the radial coordinate $r$ as $r=R(t_{0},r)$. From regularity of the center, we obtain, for $t_{0}\le t< 0$, \begin{eqnarray} F(r)&=&F_{3}r^3+\cdots,\\ R(t,r)&=&R_{1}(t)r+\cdots, \\ \nu(t,r)&=&\nu_{0}(t)+\cdots, \end{eqnarray} where ``$\cdots$'' means the higher order terms with respect to $r$. As we assume the positivity of the energy density at the center at $t=t_{0}$, $F_{3}>0$ must be satisfied. We set $\nu_{0}(t)=0$ by using the scaling freedom of time coordinate. From this choice, the time coordinate $t$ can coincide with the proper time $\tau$ at the center. Substituting into Eq.~(\ref{eq:R44timelike}), the value of $R_{44}$ at the center is written as \begin{equation} R_{44}=\frac{3 F_{3}}{R_{1}^3}. \label{eq:R44center} \end{equation} Note that $R_{1}=0$ corresponds to the occurrence of the central singularity. From the equation of motion of each mass shell (\ref{eq:energyofparticle}), it is required that \begin{equation} \frac{1}{h^2}-1=h_{1}(t)r^2+\cdots. \end{equation} The lowest order of Eq.~(\ref{eq:energyofparticle}) becomes \begin{equation} \dot{R_{1}}^2=\frac{2F_{3}}{R_{1}}+h_{1}. \label{eq:R1} \end{equation} Here we assume that GDC is satisfied for the timelike geodesic $r=0$, where it should be noted that the value of \begin{equation} \frac{R}{2m}\left(\frac{1}{h^2}-1\right) \end{equation} at $r=0$ is understood as the limit of $r\to 0$. Then it is found that \begin{equation} \lim_{t \to 0}\frac{R_{1}h_{1}}{F_{3}} =0. \end{equation} Hence, Eq.~(\ref{eq:R1}) becomes \begin{equation} \dot{R_{1}}^2\approx \frac{2F_{3}}{R_1} \end{equation} in the limit of $t\to 0$. This is integrable as \begin{equation} R_{1}\approx \left(\frac{9F_{3}}{2}\right)^{1/3}(-t)^{2/3} =\left(\frac{9F_{3}}{2}\right)^{1/3}{(-\tau)^{2/3}}. \end{equation} Eq.~(\ref{eq:R44center}) becomes \begin{equation} R_{44}\approx \frac{2}{3}\frac{1}{(-\tau)^2}. \end{equation} Therefore, for the timelike geodesic $r=0$, both SCC and LFC are satisfied. \begin{th} If GDC is satisfied for the timelike geodesic $r=0$ which terminates at the singularity, then both SCC and LFC are satisfied for the timelike geodesic. \end{th} \section{examples} \subsection{dust collapse} The spherically symmetric dust collapse has been analyzed in the context of naked singularities by Eardley and Smarr (1979), Christodoulou (1984), Newman (1986), Joshi and Dwivedi (1993), Singh and Joshi (1996), and Jhingan, Joshi and Singh (1996). The stability of the Cauchy horizon against nonspherical perturbation was recently discussed by Iguchi, Nakao and Harada (1998) and Iguchi, Harada and Nakao (1998). The dust fluid is given by $h(r,R)=h(r)$. For simplicity we restrict our attention to the marginally bound collapse which is given by $h=1$. It is trivial that the singularity is gravity-dominated. The space-time is given by the LTB solution. The solution in the mass-area coordinates is given by Ori (1990) and Magli (1998). The solution contains an arbitrary function $F(r)$. Here we choose the comoving radial coordinate $r$ as $ r=R(t=t_{0},r)$, i.e., $R^{0}(m)=F^{-1}(m)$. First we give the function $F(r)$ as \begin{equation} F(r)=F_{3}r^3+F_{5}r^5+\cdots, \label{eq:smooth} \end{equation} which corresponds to generic smooth initial data. For $F_3>0$ and $F_5<0$, Eq.~(\ref{eq:rooteq}) has a finite positive root \begin{equation} y_{0}=\left(\frac{-F_5}{4\sqrt{2}F_3^{13/6}}\right)^{2/3} \end{equation} with $\beta=7/9$. From the results in Sec.~\ref{sec:gravitydom}, not SCC but only LFC is satisfied for the radial null geodesic which emanates from the singularity. Next, if we give $F(r)$ as \begin{equation} F(r)=F_{3}r^3+F_{6}r^6+\cdots, \end{equation} which corresponds to nongeneric regular initial data. For $F_{3}>0$ and $F_{6}<-(26\sqrt{2}+15\sqrt{6})F_{3}^{5/2}$, Eq.~(\ref{eq:rooteq}) has a finite positive root $y_{0}$ with $\beta=1$, where $y_{0}> 1$ is expressed using the root of some quartic equation. Then, from the results in Sec.~\ref{sec:gravitydom}, both SCC and LFC are satisfied for the outgoing radial null geodesic which emanates from the singularity. For the above two cases, the curvature strength is exactly the same for the ingoing radial null geodesic which terminates at the singularity. On the other hand, both SCC and LFC are satisfied for the timelike geodesic $r=0$. This fact was already shown by Deshingkar, Joshi and Dwivedi (1999). This can be confirmed by the result of Sec. V since GDC is also satisfied for the timelike geodesic. This is the case not only for marginally bound collapse but also for nonmarginally bound collapse because GDC is satisfied for the timelike geodesic. \subsection{cluster of counterrotating particles} The dynamical spherical cluster of counterrotating particles was introduced and analyzed by Datta (1970), Bondi (1971) and Evans (1976). The explicit solution for the metric functions was derived by Harada, Iguchi and Nakao (1998). They also examined the occurrence of naked singularity. We again restrict our attention to the marginally bound collapse. Then, the model is given by \begin{equation} h^2=1+\frac{L^2}{R^2}, \end{equation} where $L=L(m)$ is the specific angular momentum. We give $F(r)$ as in Eq.~(\ref{eq:smooth}). If $L(m)$ is given by $L=4m$, the metric functions are expressed by elementary functions. For this case, Harada, Iguchi and Nakao (1998) showed that Eq.~(\ref{eq:rooteq}) has a finite positive root \begin{equation} y_{0}=\left(\frac{24F_3^2-F_5}{4\sqrt{2} F_3^{13/6}}\right)^{2/3}, \end{equation} for $F_5 < 24 F_3^2$ with $\beta=7/9$. Note that $F_5 < 24 F_3^2$ is the same as the requirement of no shell-crossing singularity. GDC is satisfied for the null geodesic. From the results in Sec.~\ref{sec:gravitydom}, not SCC but only LFC is satisfied for the radial null geodesic which emanates from or terminates at the singularity. On the other hand, it is found that GDC is also satisfied for the timelike geodesic $r=0$. From the result of Sec. V, both SCC and LFC are satisfied for this timelike geodesic. This is the case not only for marginally bound collapse but also for nonmarginally bound collapse because GDC is satisfied for the timelike geodesic. \subsection{\protect$\Pi=k\epsilon$} We consider the equation of state \begin{equation} \Pi=k\epsilon, \end{equation} where $k$ is a constant. This will be the simplest nontrivial equation of state for tangential pressure. Singh and Witten (1997) examined the motion of a fluid with this equation of state. From Eqs.~(\ref{eq:lapse}) and (\ref{eq:eos}), \begin{equation} \nu^{\prime}=2k\frac{R^{\prime}}{R} \end{equation} holds. Since regularity requires \begin{eqnarray} \nu&=&\nu_{0}(t)+O(r^2), \\ R&=&R_{1}(t)r+O(r^3), \end{eqnarray} it is impossible to set regular initial data for $k\ne 0$. Therefore this model is not appropriate for a probe of CCH. \section{concluding remarks} The nakedness and curvature strength of shell-focusing singularity in the spherically symmetric collapse of a fluid with vanishing radial pressure has been investigated. Along the first radial null ray from the naked singularity, $R\approx 2y_{0} m^{\beta}$ $(1/3<\beta\le 1)$ is satisfied. The $y_{0}$ and $\beta$ are determined by some root equation. The $\beta$ is closely related to the curvature strength of the singularity for the null geodesic which emanates from or terminates at the singularity. Roughly speaking, $\beta=1$ means SCC and vice versa. Then, we have defined GDC for the geodesic which emanates from or terminates at the singularity. Suppose that GDC is satisfied for the null geodesic. For this class of noncritically naked singularities, if and only if $1/3<\beta<1/2$ is satisfied, neither SCC nor LFC holds, if and only if $1/2\le \beta<1$ is satisfied, not SCC but only LFC holds, and if and only if $\beta=1$ is satisfied, both SCC and LFC hold. Furthermore, for this class of noncritically naked singularities, if and only if $\lim_{m\to 0}\lambda^{-1}R$ diverges, SCC is satisfied. We also have examined whether or not the curvature divergence condition is satisfied for a timelike geodesic. Suppose that GDC is satisfied for the timelike geodesic $r=0$ which terminates at the singularity. Then, we have found that both SCC and LFC are satisfied for the timelike geodesic. We have applied this formalism to the dust collapse and the collapse of counterrotating particles. It is noted that, with vanishing radial pressure, only if the ratio of the tangential pressure to the energy density vanishes at the center, it is possible to set regular initial data which is important ingredient when we consider physical situations. Even if we include the tangential pressure, nakedness and curvature strength of the singularity are very similar to those of the dust model if the singularity is gravity-dominated. On the other hand, if the singularity is not gravity-dominated, then we may expect that the tangential pressure plays a crucial role in the nakedness of the singularity and the extendibility of the space-time beyond the singularity. \acknowledgments We are grateful to T. Nakamura and M. Siino for helpful discussions. We are also grateful to H. Sato for his continuous encouragement. This work was partly supported by the Grant-in-Aid for Scientific Research (No. 9204) and for Creative Basic Research (No. 09NP0801) from the Japanese Ministry of Education, Science, Sports and Culture.
\section{Introduction} The brain receives an enormous amount of information transduced by peripheral sense organs. This massive information influx is coded and decoded in ways that are not yet fully understood in cognitive neuroscience. Recent studies have focussed on specific neural substrates for binding mechanisms. Binding is the process by which the brain combines different aspects of sensory modalities of one object into one unified percept. The neural mechanisms that underlie synchronization in different parts of the brain are only partly understood. There is the suggestion that synchronization may be relevant to binding \cite{binding3}. Recent experiments have shown that inhibitory interneurons in the hippocampus \cite{Whittington}, the thalamic reticular nucleus \cite{Steriade93}, and the locust olfactory system \cite{Laurent96a} can indeed synchronize neuronal discharges. Subsequent theoretical analysis of networks of interneurons has shown that strong synchronization by mutual inhibition is only moderately robust against neuronal heterogeneities \cite{wang} and synaptic noise \cite{CNS}. In strong synchronization all the neurons fire with a short time-interval from each other. In most experiments to date one measures the activity of one neuron, or a small population of neurons. Periodic oscillations (extracellular, or subthreshold intracellular) measured in these experiments are consistent with strong as well as weak synchronization. In weak synchronization the {\em average} neuronal activity is periodic, without each individual neuron having to fire at each period. Often theoretical analyses, however, have focussed on strong synchronization. Here we conjecture that weak synchronization is robust against neuronal heterogeneities and synaptic noise, and consequently it is much more likely to occur in neuronal systems. Furthermore we show that it can encode more information compared to strongly synchronized states. We present numerical results of weak synchronization in a simple model of a network of Thalamic neurons that supports our conjecture. We use a thalamic network, as an example, due to the wealth of modeling information that is already available. The mechanism we discuss here, however, has more general applicability. The thalamus acts as a relay for most of the sensory information that travels to cortical structures. It regulates sleep-wake cycles \cite{Steriade93} and it may be involved in early stimulus binding \cite{Sillito94}. The lateral geniculate nucleus (LGN) and thalamic reticular nucleus (TRN) that are involved in vision have been studied extensively. Neurons of the thalamus express low threshold Calcium currents \cite{Llinas84b}, and they rebound after a sustained hyperpolarization. It has been shown experimentally and in model calculations that inhibitory neurons can synchronize neuronal discharges in the thalamus \cite{wangrinzel93,wangetal95,GolombRinzel94,Krosigk93a,Bal95a} and produce traveling waves \cite{Rinzel98,Destexhe96,Destexhe96c,GolombRinzel96}. The hallmark of weak synchronization is multimodal interspike interval (ISI) histograms (ISIH). The ISI occurs only near multiples of a particular time-scale, e.g. the period of the population activity $T$. Multimodality of the ISIH has been observed in the LGN \cite{Funke96}, and it was attributed to the action of inhibitory neurons. Multimodal ISIH have also been found in model simulations of coupled inhibitory networks in the presence of noise \cite{Golomb92} and in systems exhibiting stochastic resonance (SR) due to a periodic drive \cite{Wies95}. A theoretical mechanism for autonomous stochastic resonance (ASR) was proposed in a recent paper \cite{Longtin97}. There the periodic drive was replaced by a periodic mode in an internal kinetic variable, such that the spikes ride on top of subthreshold voltage oscillations. In our work the periodic neuronal activity in the noise-driven system is self-induced by the network. This mechanism is absent in unconnected single neurons, or in a single element with autosynaptic feedback. It has been suggested that the brain may encode information through an ensemble or cluster of neurons that fire within a short time of each other \cite{Laurent97,Riehle97}. A particular neuron may be part of a cluster for a few cycles, before it joins another neuronal ensemble. This type of dynamics is very similar to the neuronal clusters that form in our model simulations described below. An important problem is how to quantify the information content of these binding-like cluster states. The Shannon entropy has been used as a measure of information content in investigations of sensory neurons in, for instance, crickets \cite{Levin96}, and flies \cite{Rieke97}. It is, nonetheless, not known how the brain processes information, and thus it is not clear whether the Shannon entropy is the correct quantity for this purpose. It does, however, provide an upper bound on the theoretical information content of spiking neurons. It also implies that noise in the nervous system contains information, and that noisy neurons are transmitting more information compared to regular noiseless spiking neurons. We emphasize that this statement is still controversial \cite{Softky93,Shadlen94,Shadlen98}, because even if the entropy measure yields consistent results in sensory systems this does not guarantee its relevance to the central nervous system. With these caveats in mind we still proceed to characterize the information content of our neural networks by calculating its well-defined Shannon entropy. One would like to calculate both the output entropy of the model system and the mutual information. The mutual information quantifies how the ensemble of outputs is related to one of the possible realizations of the input, and it involves additional averagings over a conditional probability distribution which makes it very hard to calculate. Even the simpler calculation of the Shannon entropy from its definition in terms of the spike times is a difficult calculation. Here we shall focus on the Shannon entropy of the neuronal output of a neuron as part of the complete network. We also present here some approximations that allows us to estimate the Shannon entropy using the interspike interval time series. \section{Methods} \subsection{Network model} Our single neuron model equation contains a low threshold Calcium current $I_{Ca}$, a general leak current $I_L$, a synaptic current $I_{syn}$, and a noise current $C_m\xi$, \begin{equation} C_m \frac{dV}{dt}=-I_{Ca}-I_L-I_{syn}-C_m\xi, \label{VOLTEQ} \end{equation} together with the first order (Hodgkin-Huxley type) kinetic equations for the activation $m$ and inactivation $h$ variables for $I_{Ca}$ and the synaptic variable $s$. This yields a neuronal dynamics in terms of four variables, $V$, $m$, $h$, and $s$. We have used the kinetics for $I_{Ca}$ and $I_{syn}$ as specified in \cite{Rinzel98} (a detailed description of the model is given in Appendix A). Our single neuron model captures some important features of the dynamics of thalamic neurons, in particular its post inhibitory rebound (PIR). We are presently studying a more complete model, incorporating thalamo-cortical relay neurons and GABAergic thalamic reticular neurons \cite{Destexhe96}, including all the relevant active currents \cite{Huguenard92,McCormick92}. Our preliminary results suggest that this does not change the conclusions of our discussion here. The neurons in our network are connected all-to-all by inhibitory GABAergic synapses. Previous studies \cite{Destexhe96c,GolombRinzel96} have shown that the precise spatial connectivity is important for the activity propagation. In this work, we will not consider the spatial characteristics of the neuronal activity. We have studied different sized systems, varying from $N=1$ (a single neuron with autosynaptic feedback) to $N=1000$. We also have included two types of noises in our model, either Gaussian current noise, characterized by $\langle \xi \rangle=0$ and \begin{equation} \langle \xi(t) \xi(0) \rangle=2D\delta(t), \end{equation} with D the strength of the noise, or with Poisson distributed excitatory post-synaptic potentials (EPSPs) and inhibitory post-synaptic potentials (IPSPs). In our previous work we have shown that these two types of noises are not fully equivalent \cite{CNS98}. Both can generate, however, similar statistics, and theoretically Gaussian noise is easier to control and vary. The results presented here are thus obtained with Gaussian noise. Unless stated differently the physiological total synaptic conductance used is $g_s=2~\mbox{mS/cm}^2$, and the decay time of the synaptic channel $\tau_s=16~\mbox{ms}$. The noise strength D is expressed in units of $mV^2/ms$, time in ms, currents in $\mu A/cm^2$, and voltage in mV. The resulting differential equations for $V_i$, $m_i$, $h_i$, $s_i$ are numerically integrated using a noise-adapted second order Runge-Kutta algorithm \cite{CNS} with a time-step $dt=0.1~\mbox{ms}$. The calculation starts with random initial conditions, with the initial voltage chosen from a uniform distribution with a range of $20~\mbox{mV}$ centered around $-68~\mbox{mV}$, and $m$, $h$, and $s$ are set equal to their asymptotic values for a given value of $V$. \subsection{Calculated quantities} The raw model output are the time-traces for $V_i$, $m_i$, $h_i$, and $s_i$. The spike-times are defined as the time when the voltage $V_i$ crosses $-30~\mbox{mV}$ from below. We determined the standard histograms of interspike intervals \cite{Gerstein62}. The instantaneous firing rate, or frequency f, is defined as the number of action potentials per second in a bin of $2~\mbox{ms}$. Both the ISI histogram and $f$ are averaged over all neurons in the network. We also calculated \begin{equation} v_{syn}=\frac{1}{N}\sum_i s_i, \label{VSYN} \end{equation} which is proportional to the current drive due to the synaptic connections with other neurons (and itself). Because the network is connected all-to-all, $v_{syn}$ is the same for each neuron and represents an averaged or mean field type drive. The variable $h$ determines the excitability of the neuron, and the state of the network strongly depends on the $h$-value distribution. When $v_{syn}$ is below a certain value the network is disinhibited and will fire shortly. To determine the $h$-distribution prior to firing we have chosen $v_{syn}=0.01$ as the threshold. This value is reached every cycle, except in the presence of strong noise. In that case the disperse nature of the firing creates an average value of $v_{syn}$ above $0.01$. We have determined both the instantaneous as well as the time-averaged distribution of $h$. \section{Results} The model considered here contains an inward low threshold Calcium current, $I_T$, that initiates the Calcium action potentials. It is inactivated at the resting membrane potential (RMP, equal to $-65.57~\mbox{mV}$), and it is de-inactivated at hyperpolarized voltages. For the neuron to be excitable, $h$ has to be de-inactivated (i.e. $h>0.305$). We illustrate this in Fig.\ref{Fig1}(a). There are two V null-clines drawn, one (I) in the absence of a current, and the other (II) in the presence of a constant hyperpolarizing current $i=-1~\mu A/cm^2$. We apply a short (10ms) and a long pulse (200ms) with strength $i$. The phase point moves to II, and the $h$ value starts increasing with time-scale $\tau_1=500~\mbox{ms}$ (see Appendix A). Upon termination of the short pulse the phase point moves back to I, without generating an action potential (AP). When the long pulse ends, however, the $h$-value is too high, and the phase point misses the nearby branch of I, and generates an AP. The necessary hyperpolarization is supplied by the inhibitory postsynaptic potential (IPSP) generated by the activity of other neurons in the network. The strength of the IPSPs is determined by the value of the synaptic conductance $g_s$ and the decay time $\tau_s$. The critical value for periodic oscillations is defined as $\tau_s=\tau_c(g_s,N)$, that depends on the number $N$ of neurons in the network. For $\tau_s<\tau_c$ the oscillation dies out after a finite number of action potentials. In Fig.$~$\ref{Fig1}(b) we show the voltage trace oscillations below and above threshold for $N=1$ (single neuron with autosynaptic feedback). At the start of the simulation the neuron is released from a hyperpolarized voltage. For subthreshold values of $\tau_s$ the neuron produces a few spikes before returning to RMP. During each spike the average $h$ value decreases, since the inhibitory drive is not strong enough to replenish the loss due to the depolarization of the AP. Above threshold a periodic spike train is produced. The $h$-value varies periodically, it decreases during the AP and it increases during the inhibition. Note that the interspike intervals are determined by $\tau_s$. The subthreshold spike train has therefore a much smaller ISI compared to the one above threshold. We have determined the boundary between stable and unstable oscillations as function of $g_s$ (Fig.$~$\ref{Fig1}(a)). The minimum duration $\tau_s$ of the IPSP needed for deinactivation, increases for weaker synaptic coupling $g_s$. The situation for a real network ($N>1$) is more complicated, since the initial voltages play an important role. If, for instance, we would start with neurons clamped at their resting membrane potential nothing will happen. To obtain a spiking network state we therefore always start the simulations with part or all the neurons clamped at hyperpolarizing voltages. With uniform initial conditions all neurons are clamped at the same voltage value. The threshold $\tau_c$ for self-sustained oscillations is then equal to the one for a single neuron (Fig.\ref{Fig2}(a)). Above threshold this network is in a coherent state: all neurons spike at the same time. For random initial conditions the initial voltage is chosen from a uniform distribution with a range of $20~\mbox{mV}$ around a hyperpolarized average value. In that case the network can sustain stable oscillations for lower values of $\tau_s$ (Fig.\ref{Fig2}(b)). The network settles in a state where groups of neurons fire simultaneously. It is easy to understand why such cluster states emerge. Starting from random initial conditions each neuron will have a different phase, and will thus reach the AP threshold at a different time. The first neurons to fire will cause an inhibition that blocks other neurons (further from threshold) from firing. They can only fire after the decay of the inhibition (a few $\tau_s$). Periodic oscillations in the network are sustained by inhibition waves produced by the activity of de-inactivated neurons. The oscillations automatically become coherent, with the initial phase differences between cluster neurons driven to zero. In the simplest cluster state each neuron fires with the same ISI. The time between consecutive cluster firings, or cycle length, may vary since the strength of inhibition depends on the cluster size. Neurons will only fire when $v_{syn}$ is below a certain value: the higher the initial value (proportional to cluster size), the longer it takes to reach this value. More complex cluster states may also contain neurons that fire with different frequencies. Sufficiently strong noise can induce and maintain a spiking network state even for $\tau_s<\tau_c$. Again we have to distinguish between single neurons and a network. The noise-induced dynamics of a single neuron with autosynaptic feedback will not yield a periodic spike-trace. Instead, the ISI distribution has a peak for short times due to ISIs within the spike trains, and an exponential distribution for the intervals between the end of one, and the start of another spike train. We show some representative voltage traces in Fig.$~$\ref{Fig3}(c). Close to threshold and with weak noise the neuron produces a long transient that dies out eventually. Stronger noise can spontaneously induce a spike. The inhibition induced by the AP then manages to produce a short spike train. The number of spikes in this train depends on the distance from threshold. This makes the dynamics discussed here essentially different from Stochastic Resonance, since in that case one would obtain a multimodal distribution, with peaks at multiples of the driving frequency. We studied the dynamics of the cluster states in a network with $N=1000$, in terms of the variables shown in Figs.$~$\ref{Fig4} and $~$\ref{Fig5}. We focused on four values for $D=0$, $0.0024$, $0.008$, and $0.8$. An important variable in our analysis is $h$, the inactivation variable of $I_T$, since the model neuron is only excitable when $h$ is large enough. The distribution of $h$ values in the network will tell us which neurons are excitable, and which ones need to be de-inactivated by further inhibition cycles. The regularity of the network dynamics is further reflected in the periodicity of $f$ and $v_{syn}$ (Fig.$~$\ref{Fig5}) and their autocorrelation function (not shown). Note that $v_{syn}$ itself acts as a drive on the neurons. The larger the distance between the peak and trough, the stronger the synchronizing force. We can identify four different types of regimes. For zero noise (D=0) the system is in a state with five clusters of unequal size. The distribution of cluster sizes is determined by the initial conditions. At each time a neuron can only have one of five h-values. This set of h-values goes through a modulation with a period of five cycles (Fig$~$\ref{Fig4}(a)). The derived quantities V, h, $v_{syn}$, and f (Fig.$~$\ref{Fig5}(a)) go through the same modulations. When all the clusters have the same size, there are no such modulations, and the $h$ histogram would consist of only $5$ peaks. Immediately after firing, the neuron is partially de-inactivated with each subsequent cycle until it is excitable again (see the $h$ time traces in Fig.$~$\ref{Fig5}). The neuron then has to wait its turn to become disinhibited and fire before other clusters do. When there is noise, there is dispersion in the spike firing times. In the absence of time delay there is only a short time interval during which neurons can fire before the inhibition blocks all other firings during one cycle. As a result, clusters lose neurons whose AP has been noise delayed, and other clusters gain those neurons as members. In addition, noise can cause neurons to fire before the rest in their cluster. Weak noise (D=0.0024) disorders the system. The network starts out with unequal cluster sizes (due to the initial conditions). Large clusters lose more neuron members than smaller ones. Weak noise, however, is not strong enough to equalize their numbers on the time-scale considered ($5\,\times10^4~\mbox{ms}$). Instead the cluster sizes start to vary in a somewhat stochastic fashion, leading to an erratic firing rate (Fig.$~$\ref{Fig4}(b)) and a fluctuating period. The time-averaged $h$ histogram is very broad (Fig.$~$\ref{Fig5}(b)). For stronger noise (D=0.008) the average cluster size becomes stationary after a brief transient. The $h$-values that the neurons of different clusters cycle go through are, on the average, the same. As a result there are six smooth peaks in the $h$-histogram. The peaks become sharper for more de-inactivated values, and the h-distribution is stationary (on the average it is the same for each cycle). The actual neurons that fire in each cluster changes with time. This state is stable up to a noise strength of approximately D=1 (for N=1000). The cycle-to-cycle fluctuations in cluster sizes increases with increasing $D$. The inhibition that each cluster receives (proportional to $v_{syn}$) varies, and as a result the width of the peaks in the $h$ histogram increases. The cluster size also decreases with increasing $D$, and the amplitude of $v_{syn}$ oscillations also decreases. The stability of a cluster can be defined as the fraction of neurons that are still part of a given cluster the next time it fires. This is related to the number and height of the peaks in the ISIH (see Fig.$~$\ref{Fig7}(a) and (b)), and it decreases with D. The neuron spends most of its time in an excited state (flat part of $h(t)$, Fig.$~$5(d)) waiting for its noise induced AP threshold to fall in the inhibition free window. For that reason it is unlikely to fire with the same cluster as in the previous time. Finally, for larger noise strengths, $D>1$, the firing is no longer organized in clusters, since the noise has become so large that during an inhibition free window not enough neurons fire coherently to create a large enough inhibition to block the discharge until the next cycle. As a result there are no quiescent periods defining cycles and no distinct cycles either. The cluster states can be quantitatively described by the average cycle length (period), and the periodicity (number of cycles) of the response. We have studied these two quantities as a function of D for different values of $\tau_s$ for N=1000, and for different system sizes for $\tau_s=16$. For smaller $\tau_s$ the amount of inhibition available for de-inactivation is lower, and the periodicity increases since the de-inactivation has to be spread out over more cycles. The cycle-period, being proportional to $\tau_s$, also decreases. In addition, the oscillation needs a higher minimum value of $D$ to sustain itself, since the average cluster size is given by the system size over the periodicity. When it becomes too small, the periodic component in the inhibition becomes too small, and as a result the cluster state dies. The noise strength for which this happens is only weakly dependent on $\tau_s$. Note that the periodic component is proportional to the cluster size normalized by the network size, in addition it decreases with increasing jitter. There is, however, a difference between large networks ($N\sim1000$) and small ones ($N\sim 10$). For small networks, a single neuron can provide enough inhibition to block neuronal discharge and thus to maintain a periodic network state. The state is very robust against noise, even when noise reduces the cluster to its minimum size (one neuron), it can still maintain a spiking state. The drawback is that the fluctuations in cluster size will be of the order of the cluster size itself, causing the cycle length to vary considerably. Below threshold (not shown) the amount of noise needed to induce a spiking state increases, and for very small networks ($N<4$) failure can occur, i.e. the network becomes quiescent if one neuron fails to fire. Note that the maximum periodicity that the system can sustain is bounded by the system size. In Fig.$~$\ref{Fig6} we see that the periodicity obtained for a given $D$ decreases with $N$. We now discuss the possible information content of the ISI time series of individual neurons. We assume that the states are characterized by a periodic population activity. This is reflected in the ISI time series because the ISI will only take values close to multiples of the cycle length. The ISIH will therefore consist of a series of peaks. If a neuron would consistently spike with the same cluster, there would be only one peak. The peak with maximum weight is close to the average ISI (and periodicity) of the network, the other peaks correspond to the ISI where the neuron changed cluster. The relative weight of the maximum peak is thus a measure of the stability of the clusters. In Figs.$~$\ref{Fig7}(a), and \ref{Fig7}(b) we show the ISIH of a state with stable and unstable clusters, respectively. The width of each peak represents the jitter around a multiple of $T$. We find that the ISIH of an individual network neuron is to a very good approximation the same as the population averaged ISIH. This does not mean that all the neurons fire independently, it only states that all individual neurons have identical properties, and that in the long run the statistics of their time-series are the same. Our analysis is therefore performed on the population averaged ISIH, and we also use the population averaged return map. We find that the ISIH is well described by a sum of Gaussians (SOG) of different widths (see Appendix B), and that the width increases with the the peak number. In Appendix B we derive an expression given in Eq.$~$(\ref{SOGentropy}) for the entropy of the SOG. We see that the number of peaks and their weight --cluster hopping-- and their width --jitter within the cluster-- yield two distinct contributions to the information encoding capacity. In Fig.$~$\ref{Fig7}(c) we plot the entropy as a function of width $\sigma$ (taken constant for all peaks) and the total number of $n$ peaks (weighted with a cosine envelope, Eq.$~$(\ref{CosEnv}). In our simulations we find that the amount of jitter and the number of peaks are closely correlated, and grow with $D$. Our present entropy calculations assume that there is no correlation between consecutive ISIs. For a given amount of correlation $\gamma$ (see Appendix B) the entropy per spike will decrease. We have quantified this in a model calculation for $n=2$ and various amounts of correlation $\gamma$ (see Fig.$~$\ref{Fig7}(d)). \section{Discussion} In recent years considerable attention, as well as controversy, has been directed at studying the variability of neuronal discharge in the cortex \cite{Shadlen98}. It is beyond doubt that neurons {\em in vivo} are noisy. The question is whether exact spike-times matter -- that is, if the jitter in spike times represents information, for instance quantified by the Shannon entropy -- or if only the average firing rate matters. If spike times {\em do} matter, then the synchronized discharge has a special significance. An important question is whether the nervous system is sensitive to synchronization or not. Our work is relevant in shedding light to this fundamental question in two ways. We have shown that noisy neurons can synchronize without the need of a strong external drive, and that the synchronized neuronal discharge has a potentially high information content. To place our results in a proper context we will now discuss these points in more detail below. The role that inhibitory interneurons play in the functioning of the nervous system has long been unclear. It is hard to find, and to record from interneurons, because they are rather small. Moreover the output of the nervous system is mostly generated by the principal neurons, so early investigations studied mainly the pyramidal neurons. In recent years it has become clear that inhibition plays a major role in synchronizing principal neurons, in for example the hippocampus \cite{traub}, the thalamus \cite{Steriade93}, and the locust olfactory system \cite{Laurent96a}. The mechanism by which synchronized oscillations are generated in the brain is only partly understood. {\em In vivo} many different rhythms of different frequency have been observed. Pharmacological manipulations of slices in {\em in vitro} experiments have elucidated some aspects of the synchronization mechanism. For instance, Whittington {\em et al} \cite{traub} showed that GABA$_A$-mediated inhibition is responsible for the synchronization in hippocampal slices. Slice experiments, however, suffer from the drawback that the natural afferents are cut, and therefore the synaptic activity giving rise to the {\em in vivo} variability is absent. Recent theoretical work has shown that in fact synchronization by mutual inhibition is not robust against neuronal heterogeneities \cite{wang} and synaptic noise \cite{CNS}. In theoretical investigations one usually only considers strong synchronization. In strong synchronization one imposes the strong constraint that each neuron has to fire within a short interval from each other. Here we propose that weak synchronization may in fact be more prevalent in networks connected by chemical synapses. In weak synchronization the {\em average} neuronal activity is periodic, without each individual neuron having to fire at each period. Weak synchronization is for example consistent with the experiments in \cite{traub}. There are exceptions. For example, recent experiments on weakly electric fish show that neurons in the pacemaker nucleus are strongly synchronized \cite{Moortgat98}. There, however, synchronization can possibly be attributed to electric gap junctions. Weak synchronization is associated with a periodic drive. This drive is either generated externally, as is the case with Stochastic Resonance \cite{Wies95,Gluck96}, or it is generated intrinsically by the network as it happens here. The neuron then skips periods, which it can either do deterministically (usually one peak in ISIH), or stochastically (multimodal ISIH). In both cases the network dynamics consists of clusters of neurons firing together. But in the latter case the neuronal composition of the clusters varies with time. A neuron that at a certain point fired with a certain cluster A, can fire the next time with another cluster B. This mechanism yields a synchronization that is robust against noise, and neuronal heterogeneity. Oscillating neural assemblies do play an important role in the functioning of the invertebrate nervous system. In a series of seminal experiments on the bee and locust olfactory system, Laurent et al. have shown that different odors activate overlapping ensembles of projection neurons \cite{Laurent94}. The periodic discharge of the ensembles is coherent on a cycle-by-cycle basis. Odors may be classified from the temporal firing pattern of projection neurons \cite{Laurent96b}. Synchronization of the projection neurons may be abolished by applying picrotoxin, and without changing their individual response characteristics \cite{Laurent96a}. This desynchronization was shown to impair the ability of bees to distinguish two closely related odors \cite{Laurent97}, and subsequently a population of neurons was found that was sensitive to the synchronization of projection neurons \cite{Laurent98}. The ability to distinguish between two odors, based on their spike trains, was then reduced under desynchronized conditions. We find that the information content, defined by the Shannon entropy of the spike-time distribution, contains three contributions. First, the jitter in the spike times around the cluster firing time. Second, the distribution of the number of cycles between two consecutive spikes, and finally the correlation between consecutive ISIs. Our analysis has been performed on the (average) output of a single neuron. One has to await the development of fast computational techniques to tackle the more challenging problem of quantifying the information output of the network, while taking into account the correlation in spike times between different neurons due to the cluster state. Our single-neuron spike-train results, however, may have direct relevance to recent experimental work on striatal neurons \cite{Wilson98}. Striatal neurons can be in the down (hyperpolarized, quiescent), or in the up-state (depolarized, noisy). The transitions to the up-state are precisely timed and synchronous. The fine-structure in the spike train is asynchronous. Wilson {\em et al.} \cite{Wilson98} have suggested that the brain may use these two channels to encode different types of information. We have studied the dependence of these oscillations on network parameters. We find that there is difference between small (around ten neurons) and large networks (a few hundred neurons), under the conditions of having a fixed total synaptic drive per neuron. For small networks one needs more noise to drive the subthreshold network into stable oscillations. These oscillations are very robust against increases in the noise level, and the fluctuations in the time between two cluster firings (cycle length) increases with the amount of noise. For large networks strong noise causes an instability, the stable cluster size for a given amount of noise becomes too small to inhibit out of sync neuronal discharges. For intermediate noise-strengths the neuronal dynamics self-organizes itself into a stochastically synchronized state. We also find that the farther the network is below threshold, more noise is necessary to induce a spiking state. The mechanism to create the oscillations is due to the competition between the excitatory de-inactivating, and the inhibitory effect of the synaptic drive. Each cycle will de-inactivate neurons, until they are excitable again. The neuron then has to await the decay of inhibition created by more excitable neurons. For some parameter values the latter stage is absent, and the dynamics is fully deinactivation dominated. The important time-scales in the dynamics are the deinactivation time-scale $\tau_1$ and the synaptic decay time $\tau_s$. The cycle or population period scales directly with $\tau_s$. Our results therefore predict that by pharmacologically decreasing $\tau_s$ one can increase the cycle frequency. In summary, the brain has circuitry capable synchronizing with heterogeneous components, and in the presence of noise. The spike trains of the synchronized discharge still contain information. Whether the brain utilizes this mechanism to synchronize, and more importantly whether it uses the information in the precise temporal sequence is still an open question awaiting further study. \section{Acknowledgements} This work was partially funded by the Northeastern University CIRCS fund, and the Sloan Center for Theoretical Neurobiology (PT). We thank TJ Sejnowski for useful suggestions. \newpage
\section{Introduction} One of the most exciting prospects of Coulomb blockade is the realization of a current standard based on periodically transferring single electrons through the device \cite{Gee90,Pot92}. In order to minimize higher order tunnelling processes which limit the accuracy, a chain of tunnelling junctions was used \cite{Kel96}. Quite recently, a single electron shuttle mechanism which is based on mechanical oscillations between two electrodes was suggested by Gorelik {\it et al.\/} \cite{Gor98}. The major advantage of a mechanically driven single electron transfer is that higher order processes are negligible. In the original suggestion of a mechanical single electron shuttle a metallic island is connected via organic molecules to two electrodes. One of the electrodes is positive and the other is negatively biased. For mechanically soft molecules self-excitation is expected to internally drive oscillations of the island which transfers charge at frequency $f$. Because of Coulomb blockade an integer number of electrons is loaded onto the grain close to one turning point, and the same number of electrons is unloaded close to the other. The number of electrons transferred per cycle is thus expected to be \begin{equation} \left<N\right>= 2n_{\rm max}\,, \hspace{1cm} n_{\rm max} =\left[\frac{CV}{e}+\frac 12\right]\,, \label{eq:N_average} \end{equation} where $V$ is the applied voltage and $C$ the island capacitance relative to the electrodes. The average current is given by $ \left<I\right>= ef\left<N\right>\,$. If the island is situated between the two electrodes there should be either $n_{\rm max}$ or $-n_{\rm max}$ electrons on it depending on which electrode it is coming from. A first step towards an experimental realization of a mechanical single electron shuttle was undertaken by Erbe {\it et al.\/} \cite{Erb98}. The authors designed a semiconductor device where a mechanical clapper oscillates between two electrodes - {\it i.e.\/} the oscillation is now externally driven with the average number of electrons transferred in each cycle of order $7\pm 2$ \cite{Erb98}. This experiment which works similar to a bell is still in the classical regime, {\it i.e.\/} Coulomb blockade is not relevant here. The aim in a future experimental setup with a metallic island on the tip of the clapper is to obtain an electron shuttle mechanism which is limited to single electrons by Coulomb repulsion. Since the tunnelling resistance $R$ increases exponentially with the distance $\Delta x$ between the island and the electrodes ($R=\exp(\Delta x/\lambda)$), tunnelling is suppressed exponentially when the island is in the middle between the two electrodes (both the oscillation amplitude and the distance between the two electrodes is much larger than $\lambda$). For the same reason both cotunnelling and simultaneous tunnelling between the island and the electrodes can be neglected. In order for electrons to tunnel the amplitude of the oscillation has to be large compared with the typical tunnelling distance $\lambda$. In the experiments by Erbe {\it et al.\/} the oscillation amplitude is of order 100 nm (compared to a few nanometers as suggested by Gorelik {\it et al.\/}) and thus much larger than $\lambda\,$. In our present work we investigate the accuracy of a mechanically driven single electron shuttle using a Master equation approach \cite{Lehrbuch}. The Master equation is solved numerically and also analytically in a simplified form. In particular, we calculate both the average number of electrons transferred per period which in general is different from the simple result (\ref{eq:N_average}) and the root mean square fluctuations $\Delta N \equiv\sqrt{\left<N^2\right>-\left<N\right>^2}\,$. As one of our main results, it turns out that optimum operation with $\Delta N\ll 1$ is obtained for long contact times ($t_0\gg RC$) and sufficiently low temperatures ($k_{\rm B}T\ll\frac{e^2}C$). \section{Master equation} In order to calculate the probability $p(m,t)$ to find $m$ additional electrons on the island at time $t\,$ we use a Master equation. This technique has been successfully used to describe both quantum dots and metallic islands exhibiting Coulomb blockade \cite{Gra91}-\cite{Bru94}. In our case the island oscillates mechanically which leads to time dependent transition rates $\Gamma(m,t)$ at the leads. Collecting gain and loss terms the Master equation reads \begin{eqnarray} \frac{d}{dt} p(m,t) =&-& \left[\Gamma^{(+)}_{\rm L}(m,t)+\Gamma^{(+)}_{\rm R}(m,t)+\Gamma^{(-)}_{\rm L}(m,t)+ \Gamma^{(-)}_{\rm R}(m,t)\right]p(m,t) \nonumber\nopagebreak\\ &+&\left[\Gamma^{(+)}_{\rm L}(m-1,t)+\Gamma^{(+)}_{\rm R}(m-1,t)\right]p(m-1,t)\nonumber\nopagebreak\\ &+&\left[\Gamma^{(-)}_{\rm L}(m+1,t)+\Gamma^{(-)}_{\rm R}(m+1,t)\right]p(m+1,t)\,. \label{eq:MasterEq} \end{eqnarray} In a golden rule approach the tunnelling rates are of the form \cite{Lehrbuch} \begin{equation} \Gamma = \frac 1 {e^2R} \, \frac {\Delta E}{1-\exp\left(-\frac{\Delta E}{k_{\rm B}T}\right)} \,. \end{equation} Since both $R$ and the capacitance $C$ in $\Delta E \propto \frac{e^2}{C}$ are time dependent now, with the dominant time dependence associated with the exponentially varying $R(t)$, the tunnelling rates factorize into: \begin{equation} \Gamma_{\rm R}^{(\mp)}(m,t) = g_{\rm R}(t)\,\Gamma_{{\rm R},m}^{(\mp)}(t) \,. \end{equation} Here \begin{equation} g_{\rm R}(t) = \frac{R_{\rm R}(t_{\rm max})C_{\rm R}(t_{\rm max})}{R_{\rm R}(t)C_{\rm R}(t)} \end{equation} is a strongly varying function of time, dominated by the exponential decrease of tunnelling when the island moves away from the electrode ($t_{\rm max}$ is the time where the island is at its closest point to the right electrode). The motion is taken as purely harmonic, {\it i.e.\/} $x(t)=x_{\rm max}\sin(\omega t)$. For the rates $\Gamma_{{\rm R},m}^{(\mp)}(t)$ whose time dependence is determined by that of the capacitance only, we take the standard result for tunnelling rates when a voltage of $-V/2$ is applied at the left electrode and a voltage of $V/2$ is applied at the right electrode: \begin{equation} \Gamma_{{\rm R},m}^{(\mp)}(t) = \frac 1{\tau} \frac{\pm \left(m+ \frac{C_{\rm L}(t)V}{e}\right) - \frac 12}{1-\exp\left[-\left(\pm\left(m + \frac{C_{\rm L}(t)V}{e}\right)- \frac 12\right)\frac{e^2}{C_{\rm \Sigma}(t)k_{\rm B}T}\right]} \end{equation} where $\tau = R_{\rm R}(t_{\rm max})C_{\rm R}(t_{\rm max})$ and $C_{\rm \Sigma}(t)=C_{\rm R}(t)+C_{\rm L}(t)$. For the left electrode the indices $R$ and $L$ have to be interchanged, $V$ has to be replaced by $-V$ and $x(t)$ by $-x(t)\,$. For the numerical calculations the full Master equation was used. For simplicity, we assumed that the capacitance depends linearly on separation like $C_{\rm L}(t) \propto\frac 1{x_0+x_{\rm max}+x(t)}\,$. The function $g(t)$ is of the form \begin{equation} g_{\rm R}(t) = \frac{x_0+x_{\rm max}-x(t)}{x_0}\exp\left[-\frac{x_{\rm max}-x(t)}{\lambda}\right] \end{equation} which is sharply peaked around $t_{\rm max}$. Since tunnelling through the left and right electrodes does not take place simultaneously ($x_{\rm max}\gg\lambda$), the tunnelling at the left or the right electrodes can be treated separately. For the analytic calculations we consider only one electrode: $ \Gamma^{(\pm)}_{\rm L}(m,t)=0\,, $ which is approached by an island with $n$ electrons from the left, {\it i.e.\/} we are calculating the conditional probability $p(m|n,t)$ to find $m$ electrons on the island given that initially there where $n$ electrons on the island. It has to be noted that when the island is close to the right electrode, $C_{\rm L}(t)$ changes much slower than $C_{\rm R}(t)$ - independent of the precise model for the capacitance. As long as the applied voltage $V$ is not too large ({\it i.e.\/} as long as $VC_{\rm L}(t_{\rm max})$ is of order $e$, $2e$, \ldots) $C_{\rm L}(t)$ can be replaced by $C_{\rm L}(t_{\rm max})\,$. In order to be able to solve the master equations analytically, we replace the function $g_{\rm R}(t)$ by a step function $\tilde{g}_{\rm R}(t)$ having the same height and area: \begin{equation} \tilde{g}_{\rm R}(t) \equiv \left\{ \begin{array}{r@{\quad:\quad}l} 0&t\le t_{\rm max}-t_0\\ 1 & t_{\rm max}-t_0<t<t_{\rm max}+t_0\\ 0 &t \ge t_0+t_{\rm max} \end{array} \right.\,,\hspace{1cm} t_0 \equiv \sqrt{\frac{\pi\lambda}{2x_{\rm max}}}\frac 1{\omega}\left(1+\frac {\lambda}{2x_0} \right)\,. \label{eq:gres} \end{equation} The width $2t_0$ corresponds to an effective contact time. As expected, the contact time decreases for increasing oscillation frequencies ($x_{\rm max} \gg \lambda\,$, $t_{\rm max}\gg t_0$). The solutions $p(m|n,t)$ of the Master equation can be used to calculate the probability $p_{\Delta}(m)$ that $m$ electrons are transferred: \begin{equation} p_{\Delta}(n) \equiv \sum_{k} p_{\rm i}(k)p({k-n|k},t_0+t_{\rm max})\,, \label{eq:p_D_def} \end{equation} where $p_{\rm i}(k)$ is the initial probability to find $k$ additional electrons on the island ($\sum_{k}p_{\rm i}(k)=1$). To calculate this probability, we consider a symmetric situation where the probability for the island to approach the right electrode with $n$ electrons is the same as the probability for the island to approach the left electrode with $-n$ electrons: \begin{equation} p_{\rm i}(n) = p_{\rm f}(-n)\,, \end{equation} where \begin{equation} p_{\rm f}({m})=\sum_{n} p_{\rm i}({n})p({m|n},t_0+t_{\rm max})\,. \label{eq:p_f_eq} \end{equation} This leads to a set of linear equations for $p_{\rm i}({n})$ which in principle can be solved for arbitrary values of $n_{\rm max}$ and $t_0$. However, for practical purposes, small values for $n_{\rm max}$ should be used. From $p_{\Delta}(n)$ we then calculate both $\left< N\right>$ and $\Delta N$ ($\left< N^k\right>\equiv \sum_NN^kp_{\Delta}(N)$). \section{Low temperature limit} For very low temperatures $k_{\rm B}T\ll\frac{e^2}{C}$ Coulomb blockade fixes the number of additional electrons $n$ on the island between $-n_{\rm max}$ and $n_{\rm max}$ strictly. We will focus on the behaviour of the single electron shuttle for small integer values of $n_{\rm max}$ ($\frac{C_{\rm L}(t_{\rm max})V}{e} = n_{\rm max}+\delta$; $-\frac 12\le\delta< \frac 12$). With the approximation described in eq.~(\ref{eq:gres}) the Master equation at $T=0$ simplifies to : \begin{equation} \frac{d}{dt}{p}({n-k}|n,t) = -\frac 1{\tau_{n-k}}{p}({n-k}|n,t) +\frac 1{\tau_{n-k+1}}{p}(n-k+1|n,t) \label{eq:master_tilde} \end{equation} with \begin{equation} \frac{1}{\tau_{-n_{\rm max}+\nu}} = \left\{ \begin{array}{r@{\quad:\quad}l}\frac 1{\tau}\left(\nu - \frac 12+\delta\right) &\nu>0\\ 0 & \nu\le0 \end{array}\right.\,. \end{equation} We solve eq.~(\ref{eq:master_tilde}) for $t_{\rm max}-t_0<t<t_{\rm max}+t_0\,$: \begin{equation} {p}(n-k|n,t) = \frac 1{k!}\left(\prod_{i=n-k+1}^n \frac{\tau}{\tau_{\rm i}}\right) \exp\left(-\frac{\Delta t}{\tau_{n-k}}\right) \left[1-\exp\left(-\frac{\Delta t}{\tau}\right)\right]^k \label{eq:p_T0} \end{equation} where $\Delta t\equiv t_{\rm max}-t_0+t$ ($n, n-k>-n_{\rm max}$). Eq.~(\ref{eq:p_T0}) can easily be proved by induction using the fact that $\frac 1{\tau_m}+\frac 1{\tau}=\frac 1{\tau_{m+1}}$. For $n=n_{\rm max}\,$, the normalization condition leads to the following result: ${p}(-n_{\rm max}|n,t)=1-\sum_{m=-n_{\rm max}+1}^n p(m|n,t)\,$. Together with eqs.~(\ref{eq:p_D_def}) and (\ref{eq:p_f_eq}) this can be used to calculate the current and the current fluctuations for $T=0\,$. The results obtained from the analytical expression~(\ref{eq:p_T0}) and also from the numerical solution of the full master equation are shown in fig.~\ref{fig:coulombblockade}. Coulomb blockade is clearly visible: for low voltages no electrons are transferred. Above a critical voltage of $V=\frac {e^2}{2C_{\rm L}(t_{\rm max})}$ the shuttle starts to work. For $n_{\rm max}=1$, $\delta = 0$ the analytic expressions for the average number of electrons transferred ($\left<N\right>$) and the mean square fluctuations ($(\Delta N)^2$) are comparatively simple: \begin{eqnarray} \left<N\right> &=& 2\frac{1-a^3} {\left[1+a\right]\left[1+\frac 12a+a^2\right]} \label{eq:NmeanT0} \\ (\Delta N)^2 &=& 2a(1-a)\frac{6+9a+22a^2+13a^3+10a^4}{\left[2a^2+a+2\right]^2\left[a+1\right]^2} \end{eqnarray} where $ a\equiv \exp\left(-\frac{t_0}{\tau}\right)\,. $ As expected, in the limit of very short contact times $t_0\ll \tau$ ($a\rightarrow 1$), no electrons are transferred whereas very long contact times $t_0\gg\tau$ ($a\rightarrow 0$) lead to the simple result~(\ref{eq:N_average}). The fluctuations disappear for both short and long contact times. Their maximum ($\Delta N\approx 0.686 $) is reached for $t_0\approx 0.92 \tau$ as shown in fig.~\ref{fig:NT0}. \section{Finite temperatures} For finite temperatures the number of additional electrons on the island is no longer as restricted as for $T=0$. Here, the Master equation has the more complicated form: \begin{eqnarray} \frac{\rm d}{{\rm d}t}p(-n_{\rm max}\!+\!k|n,t) = \frac {g(t)}{\tau}\frac{k-\frac {1}{2}+\delta}{\left(1-\varepsilon^{2k-1}\xi\right)} \left[\varepsilon^{2k-1}\xi p(-n_{\rm max}\!+\!k\!-\!1|n,t) -p\left(-n_{\rm max}\!+\!k|n,t\right)\right] \nonumber\\ + \frac {g(t)}{\tau} \frac{k+\frac {1}{2}+\delta}{\left(1-\varepsilon^{2k+1}\xi\right)} \left[p(-n_{\rm max}\!+\!k\!+\!1|n,t) -{\varepsilon^{2k+1}}\xi p(-n_{\rm max}\!+\!k|n,t)\right]\,, \end{eqnarray} where $\varepsilon \equiv \exp\left(-\frac{e^2}{2C_{\rm \Sigma}(t)k_{\rm B}T}\right)$ and $\xi = \varepsilon ^{2\delta}\,$. For $t_0\gg \tau$ and constant $C_{\rm \Sigma}$ ({\it i.e.\/} constant $\varepsilon$ and $\xi$) the stationary solution ($\frac{d}{dt}p(m|n,t)=0$, which is independent of $n$ ($p_{\rm f}(m)=p(m|n,t_{\rm max}+t_0)$) is given by: \begin{equation} p_{\rm f}(-n_{\rm max}+k) = \eta^{-1}\,{\varepsilon^{k^2}}\xi^k \,, \label{eq:p_fT} \end{equation} with the normalization constant $ {\eta\equiv \sum_{\nu=-\infty}^{\infty}\varepsilon^{\nu^2}\xi^{\nu}}\,.$ Together with eq.~(\ref{eq:p_D_def}) this leads to: \begin{equation} p_{\Delta}(2n_{\rm max}+k) =\frac{\xi^{-k}}{\eta^2} {\sum_{\nu=-\infty}^{\infty}\varepsilon^{\nu^2}\varepsilon^{(\nu+k)^2}}{} \,. \label{eq:p_deltaT} \end{equation} In reality however, $C_{\rm \Sigma}(t)$ is time-dependent and thus the results (\ref{eq:p_fT}) and (\ref{eq:p_deltaT}) are not exact. It turns out however, that if one replaces $C_{\rm \Sigma}(t)$ by $C_{\rm \Sigma}(t^*)$ where $t^*$ is defined by \begin{equation} \int_{t^*}^{\pi/\omega}g(t){\rm d}t = \tau \label{eq:def_t} \end{equation} one obtains good agreement between numerical results obtained by the solution of the time dependent problem and the analytic result (\ref{eq:p_fT}). (Changes which happen on time-scales much smaller than $\frac{\pi}{\omega}-t^*$, {\it i.e.\/} effective time-scales much smaller than $\tau$, are too fast to be followed, cf.\ fig.~\ref{fig:NT0}.) Fig. \ref{fig:coulombTg0} shows two steps of the symmetric resulting Coulomb staircase for several temperatures. In the middle of one of the steps of the Coulomb staircase ($\delta =0$) the average number of electrons transferred per period agrees with the simple result $ \left< N \right> = 2n_{\rm max} $ even for $k_{\rm B}T>\frac {e^2}{2C_{\rm \Sigma}}\,$. The root mean square fluctuations are approximately given by \begin{equation} \Delta N \simeq \left\{ \begin{array}{r@{\quad:\quad}l} 2\sqrt{\frac{\varepsilon}{2\varepsilon+1}} & k_{\rm B}T\ll\frac{e^2}{2C_{\rm \Sigma}}\\ \sqrt{k_{\rm B}T/\frac{e^2}{2C_{\rm \Sigma}}} & k_{\rm B}T\gg \frac{e^2}{2C_{\rm \Sigma}} \end{array} \right.\,, \hspace{1cm}\varepsilon = \exp\left(-\frac{e^2}{2C_{\rm \Sigma}(t^*)k_{\rm B}T}\right)\,. \end{equation} As expected, the fluctuations disappear exponentially for decreasing temperatures. \section{Summary} Using a simple Master equation we have derived analytic expressions for both the average current and its root mean square fluctuations in a mechanically driven single electron shuttle. We find a good agreement between our analytical results and a numerical solution of the fully time-dependent Master equation. To minimize fluctuations the shuttle should be operated under conditions where the interaction time $t_0$ is much greater than the $RC$-time $\tau$ ($t_0\approx 10 \tau)$) and the temperature $k_{\rm B}T$ is much lower than the Coulomb energy $\frac {e^2}{2C_{\rm \Sigma}}$. We would like to thank R.~Blick, A.~Erbe and N.~Chandra for useful discussions. Support by the {\it SFB 348 }of the {\it Deutsche Forschungsgemeinschaft }is greatfully acknowledged. \vskip-12pt
\section{Introduction} In a recent collaboration with W.~Fischler \cite{bfm}, we showed that the space of asymptotic directions in the moduli space of toroidally compactified M-theory had a hyperbolic metric, related to the hyperbolic structure of the $E_{10}$ duality group. We pointed out that this could have been anticipated from the hyperbolic nature of metric on moduli space in low energy SUGRA, which ultimately derives from the negative kinetic term for the conformal factor. An important consequence of this claim is that there are asymptotic regions of the moduli space which cannot be mapped onto either 11D SUGRA (on a large smooth manifold) or weakly coupled Type II string theory. These regions represent true singularities of M-theory at which no known description of the theory is applicable. Interestingly, the classical solutions of the theory all follow trajectories which interpolate between the mysterious singular region and the regions which are amenable to a semiclassical description. This introduces a natural arrow of time into the theory. We suggested that moduli were the natural semiclassical variables that define cosmological time in M-theory and that ``the Universe began'' in the mysterious singular region. We note that many of the singularities of the classical solutions {\it can} be removed by duality transformations. This makes the special nature of the singular region all the more striking\footnote{For reference, we note that there are actually two different types of singular region: neither the exterior of the light cone in the space of asymptotic directions, nor the past light cone, can be mapped into the safe domain. Classical solutions do not visit the exterior of the light cone.}. In view of the connection to the properties of the low energy SUGRA Lagrangian, we conjectured in \cite{bfm} that the same sort of hyperbolic structure would characterize moduli spaces of M-theory with less SUSY than the toroidal background. In this paper, we verify this conjecture for 11D SUGRA backgrounds of the form $K3 \times T^6$, which is the same as the moduli space of heterotic strings compactified on $T^9$. A notable difference is the absence of a completely satisfactory description of the safe domains of asymptotic moduli space. This is not surprising. The moduli space is known to have an F-theory limit in which there is no complete semiclassical description of the physics. Rather, there are different semiclassical limits valid in different regions of a large spacetime. \vspace{3mm} Another difference is the appearance of asymptotic domains with different internal symmetry groups. 11D SUGRA on $K3 \times T^3$ exhibits a $U(1)^{28}$ gauge group in four noncompact dimensions. At certain singularities, this is enhanced to a nonabelian group, but these singularities have finite codimension in the moduli space. Nonetheless, there are asymptotic limits in the full moduli space ({\it i.e.} generic asymptotic directions) in which the full heterotic symmetry group is restored. From the heterotic point of view, the singularity removing, symmetry breaking, parameters are Wilson lines on $T^9$. In the infinite (heterotic torus) volume limit, these become irrelevant. In this paper we will only describe the subspace of asymptotic moduli space with full $SO(32)$ symmetry. We will call this the HO moduli space from now on. The points of the moduli space will be parametrized by the dimensionless heterotic string coupling constant $g_{het}=\exp{p_0}$ and the radii $R_i=L_{het}\exp{p_i}$ where $i=1,\dots 10-d$ with $d$ being the number of large spacetime dimensions and $L_{het}$ denoting the heterotic string length. Throughout the paper we will neglect factors of order one. \vspace{3mm} Apart from these, more or less expected, differences, our results are quite similar to those of \cite{bfm}. The modular group of the completely compactified theory preserves a Lorentzian bilinear form with one timelike direction. The (more or less) well understood regimes correspond to the future light cone of this bilinear form, while all classical solutions interpolate between the past and future light cones. We interpret this as evidence for a new hyperbolic algebra ${\cal O}$, whose infinite momentum frame Galilean subalgebra is precisely the affine algebra $\hat{o} (8,24)$ of \cite{sentwod}-\cite{schwtwod}. This would precisely mirror the relation between $E_{10}$ and $E_9$. Recently, Ganor \cite{ori} has suggested the $DE_{18}$ Dynkin diagram as the definition of the basic algebra of toroidally compactified heterotic strings. This is indeed a hyperbolic algebra in the sense that it preserves a nondegenerate bilinear form with precisely one negative eigenvalue\footnote{Kac' definition of a hyperbolic algebra requires it to turn into an affine or finite dimensional algebra when one root of the Dynkin diagram is cut. We believe that this is too restrictive and that the name hyperbolic should be based solely on the signature of the Cartan metric. We thank O. Ganor for discussions of this point.}. \subsection{The bilinear form} We adopt the result of \cite{bfm} with a few changes in notation. First we will use $d=11-k$ instead of $k$ because now we start in ten dimensions instead of eleven. The parameter that makes the parallel between toroidal M-theory and heterotic compactifications most obvious is the number of large spacetime dimensions $d$. In \cite{bfm}, the bilinear form was \eqn{bilstart}{I=(\sum_{i=1}^{k}P_i)^2 + (d-2)\sum_{i=1}^k(P_i^2).} where $P_i$ (denoted $p_i$ in \cite{bfm}) are the logarithms of the radii in 11-dimensional Planck units. Now let us employ the last logarithm $P_k$ as the M-theoretical circle of a type IIA description. For the HE theory, which can be understood as M-theory on a line interval, we expect the same bilinear form where $P_k$ is the logarithm of the length of the Ho\hacek rava-Witten line interval. Now we convert (\ref{bilstart}) to the heterotic units according to the formulae ($k-1=10-d$) \eqn{bildva}{P_k=\frac 23 p_0,\qquad P_i=p_i-\frac 13 p_0, \quad i=1,\dots 10-d} where $p_0=\ln{g_{het}}$ and $p_i=\ln(R_i/ L_{het})$ for $i=1,\dots 10-d$. To simplify things, we use natural logarithms instead of the logarithms with a large base $t$ like in \cite{bfm}. This corresponds to a simple rescaling of $p$'s but the directions are finally the only thing that we study. In obtaining (\ref{bildva}) we have used the well-known formulae $R_{11}={L_{planck}^{eleven}} g_{het}^{2/3}$ and ${L_{planck}^{eleven}}=g_{het}^{1/3}L_{het}$. Substituing (\ref{bildva}) into (\ref{bilstart}) we obtain \eqn{biltri}{I=(-2p_0+\sum_{i=1}^{10-d}p_i)^2 +(d-2)\sum_{i=1}^{10-d}(p_i^2).} This bilinear form encodes the kinetic terms for the moduli in the $E_8\times E_8$ heterotic theory (HE) in the Einstein frame for the large coordinates. We can see very easily that (\ref{biltri}) is conserved by T-dualities. A simple T-duality (without Wilson lines) takes HE theory to HE theory with $R_1$ inverted and acts on the parameters like \eqn{tdu}{(p_0,p_1,p_2,\dots)\to(p_0-p_1,-p_1,p_2, \dots).} The change of the coupling constant keeps the effective 9-dimensional gravitational constant $g_{het}^2/R_1=g_{het}'^2/R'_1$ (in units of $L_{het}$) fixed. In any number of dimensions (\ref{tdu}) conserves the quantity \eqn{pten}{p_{10}=-2p_0+\sum_{i=1}^{10-d}p_i} and therefore also the first term in (\ref{biltri}). The second term in (\ref{biltri}) is fixed trivially since only the sign of $p_1$ was changed. Sometimes we will use $p_{10}$ instead of $p_0$ as the extra parameter apart from $p_1,\dots p_{10-d}$. In fact those two terms in (\ref{biltri}) are the only terms conserved by T-dualities and only the relative ratio between them is undetermined. However it is determined by S-dualities, which exist for $d\leq 4$. For the moment, we ask the reader to take this claim on faith. Since the HE and HO moduli spaces are the same on a torus, the same bilinear form can be viewed in the $SO(32)$ language. It takes the form (\ref{biltri}) in $SO(32)$ variables as well. Let us note also another interesting invariance of (\ref{biltri}), which is useful for the $SO(32)$ case. Let us express the parameters in the terms of the natural parameters of the S-dual type~I theory \eqn{sduone}{p_0=-q_0=-\ln(g_{type\,I}), \qquad p_i=q_i-\frac 12 q_0,\quad i=1,\dots 10-d} where $q_i=\ln(R_i/ L_{type\,I})$. We used $g_{type\,I}=1/ g_{het}$ and $L_{het}=g_{type\,I}^{1/2}L_{type\,I}$, the latter expresses that the tension of the D1-brane and the heterotic strings are equal. Substituing this into (\ref{biltri}) we get the same formula with $q$'s. \eqn{qbiltri}{I=(-2q_0+\sum_{i=1}^{10-d}q_i)^2 +(d-2)\sum_{i=1}^{10-d}(q_i^2)} \subsection{Moduli spaces and heterotic S-duality} Let us recall a few well-known facts about the moduli space of heterotic strings toroidally compactified to $d$ dimensions. For $d>4$ the moduli space is \eqn{modhet}{{\cal M}_d={\mathbb R}^+ \,\times\, (SO(26-d,10-d,{\mathbb Z}) \backslash SO(26-d,10-d,{\mathbb R}) / SO(26-d,{\mathbb R})\times SO(10-d,{\mathbb R})).} The factor ${\mathbb R}^+$ determines the coupling constant $L_{het}$. For $d=8$ the second factor can be understood as the moduli space of elliptically fibered K3's (with unit fiber volume), giving the duality with the F-theory. For $d=7$ the second factor also corresponds to the Einstein metrics on a K3 manifold with unit volume which expresses the duality with M-theory on K3. In this context, the factor ${\mathbb R}^+$ can be understood as the volume of the K3. Similarly for $d=5,6,7$ the second factor describes conformal field theory of type~II string theories on K3, the factor ${\mathbb R}^+$ is related to the type~IIA coupling constant. For $d=4$, i.e.\,compactification on $T^6$, there is a new surprise. The field strength $H_{\kappa\lambda\mu}$ of the $B$-field can be Hodge-dualized to a 1-form which is the exterior derivative of a dual 0-form potential, the axion field. The dilaton and axion are combined in the $S$-field which means that in four noncompact dimensions, toroidally compactified heterotic strings exhibit the $SL(2,{\mathbb Z})$ S-duality. \eqn{modhetfour}{{\cal M}_4=SL(2,{\mathbb Z})\backslash SL(2,{\mathbb R})/SO(2,{\mathbb R}) \,\times\, (SO(22,6,{\mathbb Z}) \backslash SO(22,6,{\mathbb R}) / SO(22)\times SO(6)).} Let us find how our parameters $p_i$ transform under S-duality. The S-duality is a kind of electromagnetic duality. Therefore an electrically charged state must be mapped to a magnetically charged state. The $U(1)$ symmetry expressing rotations of one of the six toroidal coordinate is just one of the 22 $U(1)$'s in the Cartan subalgebra of the full gauge group. It means that the electrically charged states, the momentum modes in the given direction of the six torus, must be mapped to the magnetically charged objects which are the KK-monopoles. The strings wrapped on the $T^6$ must be therefore mapped to the only remaining point-like\footnote{Macroscopic strings (and higher-dimensional objects) in $d=4$ have at least logarithmic IR divergence of the dilaton and other fields and therefore their tension becomes infinite.} BPS objects available, i.e.\,to wrapped NS5-branes. We know that NS5-branes are magnetically charged with respect to the $B$-field so this action of the electromagnetic duality should not surprise us. We find it convenient to combine this S-duality with T-dualities on all six coordinates of the torus. The combined symmetry $ST^6$ exchanges the point-like BPS objects in the following way: \begin{equation} \begin{array}{|rcl|} \hline \mbox{momentum modes} & \leftrightarrow & \mbox{wrapped NS5-branes}\\ \hline \mbox{wrapped strings} & \leftrightarrow & \mbox{KK-monopoles}\\ \hline \end{array}\label{tabulka} \end{equation} Of course, the distinguished direction inside the $T^6$ on both sides is the same. The tension of the NS5-brane is equal to $1/(g_{het}^2L_{het}^6)$. Now consider the tension of the KK-monopole. In 11 dimensions, a KK-monopole is reinterpreted as the D6-brane so its tension must be \eqn{tdsix}{T_{D6}=\frac{1}{g_{IIA}L_{IIA}^7}=\frac{R_{11}^2}{({L_{planck}^{eleven}})^9}} where we have used $g_{IIA}=R_{11}^{3/2}{L_{planck}^{eleven}}^{-3/2}$ and $L_{IIA}={L_{planck}^{eleven}}^{3/2}R_{11}^{-1/2}$ (from the tension of the fundamental string). The KK-monopole must always be a $(d-5)$-brane where $d$ is the dimension of the spacetime. Since it is a gravitational object and the dimensions along its worldvolume play no role, the tension must be always of order $(R_1)^2$ in appropriate Planck units where $R_1$ is the radius of the circle under whose $U(1)$ the monopole is magnetically charged. Namely in the case of the heterotic string in $d=4$, the KK-monopole must be another fivebrane whose tension is equal to \eqn{kkhet}{T_{KK5}=\frac{R_1^2}{({L_{planck}^{ten}})^8}= \frac{{R_1}^2}{g_{het}^2L_{het}^8}} where the denominators express the ten-dimensional Newton's constant. Knowing this, we can find the transformation laws for $p$'s with respect to the $ST^6$ symmetry. Here $V_6=R_1R_2R_3R_4R_5R_6$ denotes the volume of the six-torus. Identifying the tensions in (\ref{tabulka}) we get \eqn{vypocet}{\frac{1}{R'_1}=\frac{V_6}{g_{het}^2 R_1L_{het}^6},\qquad \frac{R'_1}{(L_{het}')^2}= \frac{V_6 R_1}{g_{het}^2L_{het}^8}} Dividing and multiplying these two equations we get respectively \eqn{podil}{\frac{R'_1}{L_{het}'}=\frac{R_1}{L_{het}},\qquad \frac{1}{L_{het}'}=\frac{V_6}{g_{het}^2L_{het}^7}.} It means that the radii of the six-torus are fixed in string units i.e.\,$p_1,\dots,p_6$ are fixed. Now it is straightforward to see that the effective four-dimensional $SO(32)$ coupling constant $g_{het}^2 L_{het}^6/V_6$ is inverted and the four-dimensional Newton's constant must remain unchanged. The induced transformation on the $p$'s is \eqn{indtr}{(p_0,p_1,\dots p_6, p_7,p_8 \dots)\to (p_0+m,p_1,\dots p_6, p_7+m,p_8+m\dots)} where $m=(p_1+p_2+p_3+p_4+p_5+p_6-2p_0)$ and the form (\ref{biltri}) can be checked to be constant. It is also easy to see that such an invariance uniquely determines the form up to an overall normalization i.e.\,it determines the relative magnitude of two terms in (\ref{biltri}). For $d=4$ this $ST^6$ symmetry can be expressed as $p_{10}\to -p_{10}$ with $p_1,\dots p_6$ fixed which gives the ${\mathbb Z}_2$ subgroup of the $SL(2,{\mathbb Z})$. For $d=3$ the transformation (\ref{indtr}) acts as $p_7\leftrightarrow p_{10}$ so $p_{10}$ becomes one of eight parameters that can be permuted with each other. It is a trivial consequence of the more general fact that in three dimensions, the dilaton-axion field unifies with the other moduli and the total space becomes \cite{senthreed} \eqn{modth}{{\cal M}_3= SO(24,8,{\mathbb Z})\,\backslash\, SO(24,8,{\mathbb R})\, /\, SO(24,{\mathbb R})\times SO(8,{\mathbb R}).} We have thus repaid our debt to the indulgent reader, and verified that the bilinear form (\ref{biltri}) is indeed invariant under the dualities of the heterotic moduli space for $d \geq 3$. For $d=2$ the bilinear form is degenerate and is the Cartan form of the affine algebra $\hat{o}(8,24)$ studied by \cite{sentwod}. For $d=1$ it is the Cartan form of $DE_{18}$ \cite{ori}. The consequences of this for the structure of the extremes of moduli space are nearly identical to those of \cite{bfm}. The major difference is our relative lack of understanding of the safe domain. We believe that this is a consequence of the existence of regimes like F-theory or 11D SUGRA on a large smooth K3 with isolated singularities, where much of the physics is accessible but there is no systematic expansion of all scattering amplitudes. In the next section we make some remarks about different extreme regions of the restricted moduli space that preserves the full $SO(32)$ symmetry. \section{Covering the $SO(32)$ moduli space} \subsection{Heterotic strings, type~I, type~IA and $d\geq 9$} One new feature of heterotic moduli spaces is the apparent possibility of having asymptotic domains with enhanced gauge symmetry. For example, if we consider the description of heterotic string theory on a torus from the usual weak coupling point of view, there are domains with asymptotically large heterotic radii and weak coupling, where the the full nonabelian rank $16$ Lie groups are restored. All other parameters are held fixed at what appears from the weak coupling point of view to be \lq\lq generic\rq\rq values. This includes Wilson lines. In the large volume limit, local physics is not sensitive to the Wilson line symmetry breaking. Now, consider the limit described by weakly coupled Type IA string theory on a large orbifold. In this limit, the theory consists of D-branes and orientifolds, placed along a line interval. There is no way to restore the $E_8\times E_8$ symmetry in this regime. Thus, even the safe domain of asymptotic moduli space appears to be divided into regimes in which different nonabelian symmetries are restored. Apart from sets of measure zero ({\it e.g.} partial decompactifications) we either have one of the full rank $16$ nonabelian groups, or no nonabelian symmetry at all. The example of F-theory tells us that the abelian portion of asymptotic moduli space has regions without a systematic semiclassical expansion. In a similar manner, consider the moduli space of the $E_8\times E_8$ heterotic strings on rectilinear tori. We have only two semiclassical descriptions with manifest $E_8\times E_8$ symmetry, namely HE strings and the Ho\hacek rava-Witten (HW) domain walls. Already for $d=9$ (and any $d<9$) we would find limits that are described neither by HE nor by HW. For example, consider a limit of M-theory on a cylinder with very large $g_{het}$ but the radius of the circle, $R$, in the domain $L_P \gg R \gg L_{het}^2/ {L_{planck}^{eleven}} $, and unbroken $E_8\times E_8$. We do not know how to describe this limit with any known semiclassical expansion. We will find that we can get a more systematic description of asymptotic domains in the HO case, and will restrict attention to that regime for the rest of this paper. \FIGURE[l]{\epsfig{file=d-nine.eps}\caption{The limits in $d=9$.}\label{ninefigure}} For $d=10$ there are only two limits. $p_0<0$ gives the heterotic strings and $p_0>0$ is the type I theory. However already for $d=9$ we have a more interesting picture analogous to the figure 1 in \cite{bfm}. Let us make a counterclockwise trip around the figure. We start at a HO point with $p_1=0$ which is a weakly coupled heterotic string theory with radii of order $L_{het}$ (therefore it is adjacent to its T-dual region). When we go around the circle, the radius and also the coupling increases and we reach the line $p_0\equiv (p_1-p_{10})/2=0$ where we must switch to the type I description. Then the radius decreases again so that we must perform a T-duality and switch to the type~IA description. This happens for $p_1-(p_0/2)=(3p_1+p_{10})/4=0$; we had to convert $R_1$ to the units of $L_{type\,I}=g_{het}^{1/2}L_{het}$. Then we go on and the coupling $g_{IA}$ and/or the size of the line interval increases. The most interesting is the final boundary given by $p_1=0$ which guarantees that each of the point of the $p$-space is covered precisely by one limit. \vspace{3mm} We can show that $p_1>0$ is precisely the condition that the dilaton in the type~IA theory is not divergent. Rou\-g\-h\-ly speaking, in units of $L_{type\,I}=L_{IA}$ the ``gravitational potential'' is linear in $x_1$ and proportional to $g_{IA}^2 / g_{IA}$. Here $g_{IA}^2$ comes from the gravitational constant and $1/g_{IA}$ comes from the tension of the D8-branes. Therefore we require not only $g_{IA}<1$ but also $g_{IA}<L_{type\,I} / R_{line\,interval}$. Performing the T-duality $L_{type\,I} / R_{line\,interval}=R_{circle} / L_{type\,I}$ and converting to $L_{het}$ the condition becomes precisely $R_{circle}>L_{het}$. In all the text we adopt (and slightly modify) the standard definition \cite{bfm} for an asymptotic description to be viable: dimensionless coupling constants should be smaller than one, but in cases without translational invariance, the dilaton should not diverge anywhere, and the sizes of the effective geometry should be greater than the appropriate typical scale (the string length for string theories or the Planck length for M-theory). It is important to realize that in the asymptotic regions we can distinguish between e.g. type~I and type~IA because their physics is different. We cannot distinguish between them in case the T-dualized circle is of order $L_{type\,I}$ but such vacua are of measure zero in our investigation and form various boundaries in the parameter space. This is the analog of the distinction we made between the IIA and IIB asymptotic toroidal moduli spaces in \cite{bfm} \subsection{Type IA${}^2$ and $d=8$} In $d=8$ we will have to use a new desciption to cover the parameter space, namely the double T-dual of type I which we call type~IA${}^2$. Generally, type~IA${}^k$ contains 16 D-$(9-k)$-branes, their images and $2^k$ orientifold $(9-k)$-planes. We find it also useful to perform heterotic T-dualities to make $p_i$ positive for $i=1,\dots, 10-d$ and sort $p$'s so that our interest is (without a loss of generality) only in configurations with \eqn{sort}{0\leq p_1\leq p_2\leq \dots \leq p_{10-d}} We need positive $p$'s for the heterotic description to be valid but such a transformation can only improve the situation also for type~I and its T-dual descriptions since when we turn $p$'s from negative to positive values, $g_{het}$ increases and therefore $g_{type\,I}$ decreases. For type~I we also need large radii. For its T-duals we need a very small string coupling and if we make a T-duality to convert $R>L_{type\,I}$ into $R< L_{type\,I}$, the coupling $g_{IA}$ still decreases; therefore it is good to have as large radii in the type~I limit as possible. \vspace{3mm} In $d=8$ our parameters are $p_0,p_1,p_2$ or $p_{10},p_1,p_2$ where $p_{10}=-2p_0+p_1+p_2$ and we will assume $0< p_1< p_2$ as we have explained (sets of measure zero such as the boundaries between regions will be neglected). If $p_0<0$, the HO description is good. Otherwise $p_0>0$. If furthermore $2p_1-p_0>0$ (and therefore also $2p_2-p_0>0$), the radii are large in the type~I units and we can use the (weakly coupled) type~I description. Otherwise $2p_1-p_0<0$. If furthermore $2p_2-p_0>0$, we can use type~IA strings. Otherwise $2p_2-p_0<0$ and the type~IA${}^2$ description is valid. Therefore we cover all the parameter space. Note that the F-theory on K3 did not appear here. In asymptotic moduli space, the F-theory regime generically has no enhanced nonabelian symmetries. \vspace{3mm} In describing the boundaries of the moduli space, we used the relations $L_{het}=g_{type\,I}^{1/2}L_{type\,I}$, $g_{het}=1/g_{type\,I}$. The condition for the dilaton not to diverge is still $p_1>0$ for any type~IA${}^k$ description. The longest direction of the $T^k/ {\mathbb Z}_2$ of this theory is still the most dangerous for the dilaton divergence and is not affected by the T-dualities on the shorter directions of the $T^k/ {\mathbb Z}_2$ orientifold. For $d=9$ (and fortunately also for $d=8$) the finiteness of the dilaton field automatically implied that $g_{IA^k} < 1$. However this is not true for general $d$. After a short chase through a sequence of S and T-dualities we find that the condition $g_{IA^k}<1$ can be written as \eqn{karkulka}{(k-2)p_0 -2\sum_{i=1}^k p_i<0.} We used the trivial requirement that the T-dualities must be performed on the shortest radii (if $R_j< L_{type\,I}$, also $R_{j-1} < L_{type\,I}$ and therefore it must be also T-dualized). Note that for $k=1$ the relation is $-p_0-2p_1<0$ which is a trivial consequence of $p_1>0$ and $p_0>0$. Also for $k=2$ we get a trivial condition $-2(p_1+p_2)<0$. However for $k>2$ this condition starts to be nontrivial. This is neccessary for consistency: otherwise IA${}^k$ theories would be sufficient to cover the whole asymptotic moduli space, and because of S-dualities we would cover the space several times. It would be also surprising not to encounter regimes described by large K3 geometries. \subsection{Type IA${}^3$, M-theory on K3 and $d=7$} This happens already for $d=7$ where the type~IA${}^3$ description must be added. The reasoning starts in the same way: for $p_0<0$ HO, for $2p_1-p_0>0$ type~I, for $2p_2-p_0>0$ type~IA, for $2p_3-p_0>0$ type~IA${}^2$. However, when we have $2p_3-p_0<0$ we cannot deduce that the conditions for type~IA${}^3$ are obeyed because also (\ref{karkulka}) must be imposed: \eqn{kremilek}{p_0-2(p_1+p_2+p_3)<0} It is easy to see that this condition is the weakest one i.e.\,that it is implied by any of the conditions $p_0<0$, $2p_1-p_0>0$, $2p_2-p_0>0$ or $2p_3-p_0>0$. Therefore the region that we have not covered yet is given by the opposite equation % \eqn{vochomurka}{2p_0-4(p_1+p_2+p_3)=-p_{10}-3(p_1+p_2+p_3)>0} % The natural hypothesis is that this part of the asymptotic parameter space is the limit where we can use the description of M-theory on a K3 manifold. However things are not so easy: the condition that $V_{K3}> ({L_{planck}^{eleven}})^4$ gives just $p_{10}<0$ which is a weaker requirement than (\ref{vochomurka}). The K3 manifold has a $D_{16}$ singularity but this is not the real source of the troubles. A more serious issue is that the various typical sizes of such a K3 are very different and we should require that each of them is greater than ${L_{planck}^{eleven}}$ (which means that the shortest one is). In an analogous situation with $T^4$ instead of K3 the condition $V_{T^4}>{L_{planck}^{eleven}}^4$ would be also insufficient: all the radii of the four-torus must be greater than ${L_{planck}^{eleven}}$. Now we would like to argue that the region defined by (\ref{vochomurka}) with our gauge $0<p_1<p_2<p_3$ can indeed be described by the 11D SUGRA on K3, except near the $D_{16}$ singularity. Therefore, all of the asymptotic moduli space is covered by regions which have a reasonable semiclassical description. While the fourth root of the volume of K3 equals \eqn{arabela}{\frac{V_{K3}^{1/4}}{{L_{planck}^{eleven}}}=\frac{g_{het}^{1/3}L_{het}^{1/2}}{V_3^{1/6}} =\exp\left(p_0/3-(p_1+p_2+p_3)/6\right)=\exp(-p_{10}/6),} the minimal typical distance in K3 must be corrected to agree with (\ref{vochomurka}). We must correct it only by a factor depending on the three radii in heterotic units (because only those are the parameters in the moduli space of metric on the K3) so the distance equals (confirming (\ref{vochomurka})) \eqn{rumburak}{\frac{L_{min.K3}}{{L_{planck}^{eleven}}}=\exp\left(-p_{10}/6-(p_1+p_2+p_3)/2 \right).} Evidence that (\ref{rumburak}) is really correct and thus that we understand the limits for $d=7$ is the following. We must first realize that 16 independent two-cycles are shrunk to zero size because of the $D_{16}$ singularity present in the K3 manifold. This singularity implies a lack of understanding of the physics in a vicinity of this point but it does not prevent us from describing the physics in the rest of K3 by 11D SUGRA. So we allow the 16 two-cycles to shrink. The remaining 6 two-cycles generate a space of signature 3+3 in the cohomology lattice: the intersection numbers are identical to the second cohomology of $T^4$. We can compute the areas of those 6 two-cycles because the M2-brane wrapped on the 6-cycles are dual to the wrapped heterotic strings and their momentum modes. Now let us imagine that the geometry of the two-cycles of K3 can be replaced by the 6 two-cycles of a $T^4$ which have the same intersection number. \vspace{3mm} It means that the areas can be written as $a_1a_2,a_1a_3,a_1a_4$, $a_2a_3,a_2a_4,a_3a_4$ where $a_1,a_2,a_3,a_4$ are the radii of the four-torus and correspond to some typical distances of the K3. If we order the $a$'s so that $a_1<a_2<a_3<a_4$, we see that the smallest of the six areas is $a_1a_2$ (the largest two-cycle is the dual $a_3a_4$) and similarly the second smallest area is $a_1a_3$ (the second largest two-cycle is the dual $a_2a_4$). On the heterotic side we have radii $L_{het}<R_1<R_2<R_3$ (thus also $L_{het}^2/R_3< L_{het}^2/R_2< L_{het}^2/R_1<L_{het}$) and therefore the correspondence between the membranes and the wrapping and momentum modes of heterotic strings tells us that \eqn{budulinek}{\frac{a_1a_2}{{L_{planck}^{eleven}}^3}=\frac{1}{R_3}, \quad \frac{a_3a_4}{{L_{planck}^{eleven}}^3}=\frac{R_3}{L_{het}^2}, \qquad \frac{a_1a_3}{{L_{planck}^{eleven}}^3}=\frac{1}{R_2}, \quad \frac{a_2a_4}{{L_{planck}^{eleven}}^3}=\frac{R_2}{L_{het}^2}.} As a check, note that $V_{K3}=a_1a_2a_3a_4$ gives us ${L_{planck}^{eleven}}^6/ L_{het}^2$ as expected (since heterotic strings are M5-branes wrapped on $K3$). We will also assume that \eqn{liska}{\frac{a_1a_4}{{L_{planck}^{eleven}}^3}=\frac{1}{R_1}, \quad \frac{a_2a_3}{{L_{planck}^{eleven}}^3}=\frac{R_1}{L_{het}^2}.} Now we can calculate the smallest typical distance on the K3. \eqn{jezinka}{a_1=\sqrt{\frac{a_1a_2\cdot a_1a_3}{a_2a_3}}= \frac{{L_{planck}^{eleven}}^{3/2}L_{het}}{\sqrt{R_1R_2R_3}}} which can be seen to coincide with (\ref{rumburak}). There is a subtlety that we should mention. It is not completely clear whether $a_1a_4<a_2a_3$ as we assumed in (\ref{liska}). The opposite possibility is obtained by exchanging $a_1a_4$ and $a_2a_3$ in (\ref{liska}) and leads to $a_1$ greater than (\ref{jezinka}) which would imply an overlap with the other regions. Therefore we believe that the calculation in (\ref{liska}) and (\ref{jezinka}) is the correct way to find the condition for the K3 manifold to be large enough for the 11-dimensional supergravity (as a limit of M-theory) to be a good description. \subsection{Type IA${}^{4,5}$, type~IIA/B on K3 and $d=6,5$} Before we will study new phenomena in lower dimensions, it is useful to note that in any dimension we add new descriptions of the physics. The last added limit always corresponds to the ``true'' S-dual of the original heterotic string theory -- defined by keeping the radii fixed in the heterotic string units (i.e.\,also keeping the shape of the K3 geometry) and sending the coupling to infinity -- because this last limit always contains the direction with $p_0$ large and positive (or $p_{10}$ large and negative) and other $p_i$'s much smaller. \begin{itemize} \item In 10 dimensions, the true S-dual of heterotic strings is the type~I theory. \item In 9 dimensions it is type~IA. \item In 8 dimensions type~IA${}^2$. \item In 7 dimensions we get M-theory on K3. \item In 6 dimensions type IIA strings on K3. \item In 5 dimensions type IIB strings on K3$\times S^1$ where the circle decompactifies as the coupling goes to infinity. The limit is therefore a six-dimensional theory. \item In 4 dimensions we observe a mirror copy of the region $p_{10}<0$ to arise for $p_{10}>0$. The strong coupling limit is the heterotic string itself. \item In 3 dimensions the dilaton-axion is already unified with the other moduli so it becomes clear that we studied an overly specialized direction in the examples above. Nevertheless the same claim as in $d=4$ can be made. \item In 2 dimensions only positive values of $p_{10}$ are possible therefore the strong coupling limit does not exist in the safe domain of moduli space. \item In 1 dimension the Lorentzian structure of the parameter space emerges. Only the future light cone corresponds to semiclassical physics which is reasonably well understood. The strong coupling limit defined above would lie inside the unphysical past light cone. \end{itemize} Now let us return to the discussion of how to separate the parameter space into regions where different semiclassical descriptions are valid. We may repeat the same inequalities as in $d=7$ to define the limits HO, I, IA, IA${}^2$, IA${}^3$. But for M-theory on K3 we must add one more condition to the constraint (\ref{vochomurka}): a new circle has been added and its size should be also greater than ${L_{planck}^{eleven}}$. For the new limit of the type~IIA strings on K3 we encounter similar problems as in the case of the M-theory on K3. Furthermore if we use the definition (\ref{rumburak}) and postulate this shortest distance to be greater than the type~IIA string length, we do not seem to get a consistent picture covering the whole moduli space. Similarly for $d=5$, there appear two new asymptotic descriptions, namely type~IA${}^5$ theory and type~IIB strings on $K3\times S^1$. It is clear that the condition $g_{IA^5}<1$ means part of the parameter space is not understood and another description, most probably type~IIB strings on $K3\times S^1$, must be used. Unfortunately at this moment we are not able to show that the condition for the IIB theory on K3 to be valid is complementary to the condition $g_{IA^5}<1$. A straightforward application of (\ref{jezinka}) already for the type~IIA theory on a K3 gives us a different inequality. Our lack of understanding of the limits for $d<7$ might be solved by employing a correct T-duality of the type~IIA on K3 but we do not have a complete and consistent picture at this time. \subsection{Type IA${}^6$ and S-duality in $d=4$} Let us turn to the questions that we understand better. As we have already said, in $d=4$ we see the ${\mathbb Z}_2$ subgroup of the $SL(2,{\mathbb Z})$ S-duality which acts as $p_{10}\to -p_{10}$ and $p_1,\dots,p_6$ fixed in our formalism. This reflection divides the $p$-space to subregions $p_{10}>0$ and $p_{10}<0$ which will be exchanged by the S-duality. This implies that a new description should require $p_{10}>0$. Fortunately this is precisely what happens: in $d=4$ we have one new limit, namely the type~IA${}^6$ strings and the condition (\ref{karkulka}) for $g_{IA^6}<1$ gives \eqn{myslivec}{4p_0-2\sum_{i=1}^6 p_i =-2p_{10}<0} or $p_{10}>0$. \vspace{3mm} In the case of $d=3$ we find also a fundamental domain that is copied several times by S-dualities. This fundamental region is again bounded by the condition $g_{gauge}^{eff.4-dim}<1$ which is the same like $g_{IA^6}<1$ and the internal structure has been partly described: the fundamental region is divided into several subregions HO, type~I, type~IA${}^k$, M/K3, IIA/K3, IIB/K3. As we have said, we do not understand the limits with a K3 geometry well enough to separate the fundamental region into the subregions enumerated above. We are not even sure whether those limits are sufficient to cover the whole parameter space. In the case of $E_8\times E_8$ theory, we are pretty sure that there are some limits that we do not understand already for $d=9$ and similar claim can be true in the case of the $SO(32)$ vacua for $d<7$. We understand much better how the entire parameter space can be divided into the copies of the fundamental region and we want to concentrate on this question. The inequality $g_{gauge}^{eff.4-dim}<1$ should hold independently of which of the six radii are chosen to be the radii of the six-torus. In other words, it must hold for the smallest radii and the condition is again (\ref{myslivec}) which can be for $d=3$ reexpressed as $p_{7}<p_{10}$. So the ``last'' limit at the boundary of the fundamental region is again type~IA${}^6$ and not type~IA${}^7$, for instance. It is easy to show that the condition $g_{IA^6}<1$ is implied by any of the conditions for the other limits so this condition is the weakest of all: all the regions are inside $g_{IA^6}<1$. This should not be surprising, since $g_{gauge}^{eff.4-dim}=(g_{IA^6})^{1/2} =g_{IA^6}^{open}$; the heterotic S-duality in this type~IA${}^6$ limit can be identified with the S-duality of the effective low-energy description of the D3-branes of the type~IA${}^6$ theory. As we have already said, this inequality reads for $d=3$ \eqn{cipisek}{2p_0-\sum_{i=1}^6 p_i =-p_{10}+p_7<0} or $p_{10}>p_7$. We know that precisely in $d=3$ the S-duality (more precisely the $ST^6$ transformation) acts as the permutation of $p_7$ and $p_{10}$. Therefore it is not hard to see what to do if we want to reach the fundamental domain: we change all signs to pluses by T-dualities and sort all {\it eight} numbers $p_1,\dots p_7; p_{10}$ in the ascending order. The inequality (\ref{cipisek}) will be then satisfied. The condition $g_{gauge}^{eff.4-dim}<1$ or (\ref{myslivec}) will define the fundamental region also for the case of one or two dimensions. \subsection{The infinite groups in $d\leq 2$} In the dimensions $d>2$ the bilinear form is positive definite and the group of dualities conserves the lattice ${\mathbb Z}^{11-d}$ in the $p$-space. Therefore the groups are finite. However for $d=2$ (and {\it a fortiori} for $d=1$ because the $d=2$ group is isomorphic to a subgroup of the $d=1$ group) the group becomes infinite. In this dimension $p_{10}$ is unchanged by T-dualities and S-dualities. The regions with $p_{10}\leq 0$ again correspond to mysterious regions where the holographic principle appears to be violated, as in \cite{bfm}. Thus we may assume that $p_{10}=1$; the overall normalization does not matter. Start for instance with $p_{10}=1$ and \eqn{rumcajs}{(p_1,p_2,\dots p_8)=(0,0,0,0,0,0,0,0)} and perform the S-duality ($ST^6$ from the formula (\ref{indtr})) with $p_7$ and $p_8$ understood as the large dimensions (and $p_1\dots p_6$ as the 6-torus). This transformation maps $p_7\mapsto p_{10}-p_8$ and $p_8\mapsto p_{10}-p_7$. So if we repeat $ST^6$ on $p_7,p_8$, T-duality of $p_7,p_8$, $ST^6$, $T^2$ and so on, $p_{1}\dots p_6$ will be still zero and the values of $p_7,p_8$ are \eqn{rakosnicek}{(p_7,p_8)=(1,1)\to(-1,-1)\to(2,2)\to(-2,-2)\to(3,3) \to\dots} and thus grow linearly to infinity, proving the infinite order of the group. The equation for $g_{IA^6}<1$ now gives \eqn{bobek}{2p_0-\sum_{i=1}^6 p_i =-p_{10}+p_7+p_8<0} or $p_{10}>p_7+p_8$. Now it is clear how to get to such a fundamental region with (\ref{bobek}) and $0<p_1<\dots p_8$. We repeat the $ST^6$ transformation with the two largest radii ($p_7,p_8$) as the large coordinates. After each step we turn the signs to $+$ by T-dualities and order $p_1<\dots< p_8$ by permutations of radii. A bilinear quantity decreases assuming $p_{10}>0$ and $p_{10}<p_7+p_8$ much like in \cite{bfm}, the case $k=9$ ($d=2$): \eqn{pokuston}{C_{d=2}=\sum_{i=1}^8 (p_i)^2\to \sum_{i=1}^8 (p_i)^2+2p_{10}(p_{10}-(p_7+p_8))} In the same way as in \cite{bfm}, starting with a rational approximation of a vector $\vec p$, the quantity $C_{d=2}$ cannot decrease indefinitely and therefore finally we must get to a point with $p_{10}>p_7+p_8$. \vspace{3mm} In the case $d=1$ the bilinear form has a Minkowski signature. The fundamental region is now limited by \eqn{hurvinek}{2p_0-\sum_{i=1}^6 p_i =-p_{10}+p_7+p_8+p_9<0} and it is easy to see that under the $ST^6$ transformation on radii $p_1\dots p_6$, $p_{10}$ transforms as \eqn{spejbl}{p_{10}\to 2p_{10}-(p_7+p_8+p_9).} Since the $ST^6$ transformation is a reflection of a spatial coordinate in all cases, it keeps us inside the future light cone if we start there. Furthermore, after each step we make such T-dualities and permutations to ensure $0<p_1<\dots p_9$. If the initial $p_{10}$ is greater than $[(p_1)^2+\dots+(p_9)^2]^{1/2}$ (and therefore positive), it remains positive and assuming $p_{10}<p_7+p_8 +p_9$, it decreases according to (\ref{spejbl}). But it cannot decrease indefinitely (if we approximate $p$'s by rational numbers or integers after a scale transformation). So at some point the assumption $p_{10}<p_7+p_8+p_9$ must break down and we reach the conclusion that fundamental domain is characterized by $p_{10}>p_7+p_8+p_9$. \subsection{The lattices} In the maximally supersymmetric case \cite{bfm}, we encountered exceptional algebras and their corresponding lattices. We were able to see some properties of the Weyl group of the exceptional algebra $E_{10}$ and define its fundamental domain in the Cartan subalgebra. In the present case with 16 supersymmetries, the structure of lattices for $d>2$ is not as rich. The dualities always map integer vectors $p_i$ onto integer vectors. For $d>4$, there are no S-dualities and our T-dualities know about the group $O(26-d,10-d,{\mathbb Z})$. For $d=4$ our group contains an extra ${\mathbb Z}_2$ factor from the single S-duality. For $d=3$ they unify to a larger group $O(8,24,{\mathbb Z})$. We have seen the semidirect product of $({\mathbb Z}_2)^8$ and $S_8$ related to its Weyl group in our formalism. For $d=2$ the equations of motion exhibit a larger affine $\hat o(8,24)$ algebra whose discrete duality group has been studied in \cite{sentwod}. In $d=1$ our bilinear form has Minkowski signature. The S-duality can be interpreted as a reflection with respect to the vector \eqn{refsl}{(p_1,p_2,\dots, p_9,p_{10})=(0,0,0,0,0,0,-1,-1,-1,+1).} This is a spatial vector with length-squared equal to minus two (the form (\ref{biltri}) has a time-like signature). As we have seen, such reflections generate together with T-dualities an infinite group which is an evidence for an underlying hyperbolic algebra analogous to $E_{10}$. Indeed, Ganor \cite{ori} has argued that the $DE_{18}$ ``hyperbolic'' algebra underlies the nonperturbative duality group of maximally compactified heterotic string theory. The Cartan algebra of this Dynkin diagram unifies the asymptotic directions which we have studied with compact internal symmetry directions. Its Cartan metric has one negative signature direction. \section{Conclusions} The parallel structure of the moduli spaces with 32 and 16 SUSYs gives us reassurance that the features uncovered in \cite{bfm} are general properties of M-theory. It would be interesting to extend these arguments to moduli spaces with less SUSY. Unfortunately, we know of no algebraic characterization of the moduli space of M-theory on a Calabi Yau threefold. Furthermore, this moduli space is no longer an orbifold. It is stratified, with moduli spaces of different dimensions connecting to each other via extremal transitions. Furthermore, in general the metric on moduli space is no longer protected by nonrenormalization theorems, and we are far from a characterization of all the extreme regions. For the case of four SUSYs the situation is even worse, for most of what we usually think of as the moduli space actually has a superpotential on it, which generically is of order the fundamental scale of the theory. \footnote{Apart from certain extreme regions, where the superpotential asymptotes to zero, the only known loci on which it vanishes are rather low dimensional subspaces of the classical moduli space, \cite{bdw}.} There are thus many hurdles to be jumped before we can claim that the concepts discussed here and in \cite{bfm} have a practical application to realistic cosmologies. \vspace{15mm} \acknowledgments We are grateful to Ori Ganor for valuable discussions. This work was supported in part by the DOE under grant number DE-FG02-96ER40559. \newpage
\section{Overview} Current observations of gamma-ray bursts place a number of strong constraints on gamma-ray burst physics. The measured red shift of $z = 0.835$, $0.966$, and $1.61$ for lines in the optical emission of gamma-ray bursts GRB~970508 (Metzger \etal\ \markcite{Metzger1}1997a,\markcite{Metzger1}b), GRB~980703 (Djorgovski \etal\ \markcite{Djorgovski1}1998a,\markcite{Djorgovski2}b), and GRB~990123 (\markcite{Kelson}Kelson \etal\ 1999), respectively, and the indirect redshifts of $z = 3.5$, $5$, and $1.096$ for GRB~971214 (\markcite{Kulkarni}Kulkarni, S. R., \etal\ 1998), GRB~980329 (\markcite{Fruchter}Fruchter 1999), and GRB~980613 (\markcite{Djorgovski3}Djorgovski \etal\ 1999), respectively, show that gamma-ray bursts are at extraordinarily-high redshifts. The power-law gamma-ray spectrum above $511 \, \keV$ and the rapid rise times of the gamma-ray light-curves force one to consider theories with highly-relativistic bulk motion in order to avoid thermalization of the gamma-rays through photon-photon pair creation (\markcite{Schmidt}Schmidt 1978; \markcite{Baring}Baring, \& Harding 1996). The gamma-ray spectrum can be characterized by the E-peak energy ($E_p$), which is the photon energy of the maximum of the $\nu F_{\nu}$ spectrum. Because the distribution of $E_p$ values is narrow, with an average value of $E_p \approx 250 \, \keV$ (\markcite{Mallozzi}Mallozzi \etal\ 1996; \markcite{Brainerd4}Brainerd \etal\ 1999), one suspects that the characteristics of the spectrum are independent of the bulk Lorentz factor, which should vary greatly from burst to burst. In some bursts, the shape of the gamma-ray spectrum is inconsistent with optically thin synchrotron emission (Preece \etal\ \markcite{Preece2}1998a,\markcite{Preece3}b). The shape of the burst spectrum is consistent with being Compton attenuated by a high density ($\approx 10^5 \, \cm^{-3}$) interstellar medium (\markcite{Brainerd1}Brainerd 1994; Brainerd \etal\ \markcite{Brainerd2}1996,\markcite{Brainerd3}1998); observations show the x-ray excess predicted by this theory (\markcite{Preece4}Preece \etal\ 1996), and the redshifts of $\approx 1$ to $\approx 10$ derived by fitting the model to burst spectra are consistent with the values measured at optical wavelength (\markcite{Preece1}Preece \& Brainerd 1999). The success of this theory, its presence in every gamma-ray burst, implies that a high density interstellar medium is necessary for a burst to occur. The observations suggest that the sources of gamma-ray bursts are compact---perhaps a several mass black hole, perhaps a supermassive black hole---and that these sources eject mass at relativistic velocities in one short event. The observed behavior of the gamma-ray burst arises from processes that convert the kinetic energy of the ejected material into electromagnetic radiation. The general view that has developed concerning the gamma-ray emission of gamma-ray bursts is that it is radiated behind a shock that has developed within a shell moving with a bulk Lorentz factor of $\Gamma \approx 10^3$ (\markcite{Meszaros}M\'esz\'aros 1998). The radiative mechanism universally cited is synchrotron emission from electrons accelerated to an energy close to the equipartition energy (\markcite{Tavani}Tavani 1996). The shock itself is collisionless, arising either when the shell runs into the interstellar medium, or when the fastest portions of the shell overtake slower portions (Rees \& Mesaros \markcite{Rees1}1992, \markcite{Rees2}1994; \markcite{Piran}Piran \& Sari 1998; \markcite{Mochkovitch}Mochkovitch \& Daigne 1998). This model has several shortcomings. First, the gamma-ray burst spectrum is often harder at x-ray energies than is allowed by an optically-thin synchrotron emission model. Second, it provides no explanation for the x-ray excess, for the narrow $E_p$ distribution, or for the apparent absence of bursts with $\Gamma < 10^3$, as inferred from the absence of photon-photon pair creation and thermalization. From the theoretical side, many assumptions are made without theoretical development; among these are that collective processes exist that mediate the shock, that a strong magnetic field is generated in the shock, and that the energy is efficiently transferred from the bulk-motion of the ions to the thermal energy of the electrons. It is with these difficulties in mind that a new theory for the generation of the prompt radiation in a gamma-ray burst is proposed. The theory is a plasma instability theory. In this theory, the shell of relativistic material that is ejected from the source interacts with the interstellar medium through the plasma filamentation instability, which, in the relativistic regime, has a higher growth rate than the two-stream instability. Because the mass ejection event must be short, of order the burst durations, and because the mass travels $2 c \Gamma^2$ times the burst duration, the interaction is of a thin shell with the interstellar medium. As the shell passes through the interstellar medium, the interstellar medium collapses into filaments that contain strong magnetic fields and high electron thermal energies. It is the interstellar medium that filaments rather than the shell, because the interstellar medium has the smaller density as measured in each comoving reference frame. The ions in the filaments remain essentially at rest with respect to the observer, while the electrons move towards the observer with a high bulk Lorentz factor. The magnetic fields generated through filamentation are strong enough to produce gamma-rays through synchrotron self-Compton emission. One finds that the theory places lower limits on both $\Gamma$ and the interstellar medium density through selection effects, and that these lower limits lead to the conditions required by the Compton attenuation theory. The theory implies that there exists a class of burst that produces intense and prompt optical and ultraviolet emission, but no x-rays and gamma-rays. In this article, I give an analytic development of the theory outlined above. This development explores the relevant plasma and radiative processes, deriving the selection effects inherent in the theory, and ascertaining the aspects of the theory that provide means of observationally testing the theory. In \S 2, the characteristic ratio of shell density to interstellar medium density and the characteristic thickness of the shell are discussed. In \S 3, the growth of the two-stream and filamentation plasma instabilities is examined. The saturation of the filaments and the inability of the filamentation instability to mediate a shock are examined in \S 4. The electron thermalization and isotropization and the rest frame defined by the electron component are discussed in \S 5. The radiative processes of synchrotron emission and synchrotron self-Compton emission are examined in \S 6, where the characteristic frequencies and emission rates of each are derived. In \S 7, the radiative timescales for synchrotron and synchrotron self-Compton cooling are derived. The theory has several natural selection effects that constrain several of the free parameters in the theory. These are discussed in \S 8. The basic theory and the results of this study are summarized in \S9. This section also contains some suggestions for observational tests of the validity of the theory. \section{Characteristic Model Parameters} The model is of a fully-ionized neutral shell of electrons and protons passing through a fully-ionized interstellar medium of electrons and protons. The characteristic physical parameters that describe this theory are $n_{ism}$, the number density of the electrons or of the protons in the interstellar material, $n_{shell}^{\prime}$, the number density of the electrons or protons in the relativistic shell, and $\Gamma$, the Lorentz factor of the shell relative to the interstellar medium. The quantities with primes are measured in the shell rest frame, while those without primes are measured in the interstellar medium rest frame. An important parameter in the discussion that follows is the parameter $\eta$, defined as \begin{equation} \eta = { n_{ism} \over n_{shell}^{\prime} } \, . \end{equation} There is no simple connection between $\eta$, $\Gamma$, and the parameters that describe the energetics of the system. In particular, to relate $\eta$ to ${\cal M}$, the mass per unit ster radian of the relativistic shell, and $R$, the distance traveled before deceleration, requires a good understanding of the radiative transfer and plasma physics of the problem. The value of $\eta$ is therefore treated as a free parameter, with the limits on the acceptable values of $\eta$ set by a combination of theoretical and observational constraints. The former is set in this section, while the latter is set in \S 8. A specific value of $\eta$ that has a physical significance is the value of $\eta$ expected from the continuity equation for a relativistic shock. For $\Gamma \gg 1$, one has $n_{shell}^{\prime} = n_{ism} \Gamma$, and \begin{equation} \eta_{shock} = { 1 \over \Gamma} \, . \end{equation} When $\eta < \eta_{shock}$, the interstellar medium density in the shell rest frame is less than the shell density, while it is greater than the shell density when $\eta > \eta_{shock}$. A lower limit $\eta_{min}$ on $\eta$ can be derived through a theoretical argument concerning pressure equilibrium with in the shell. If one assumes that the interaction with the interstellar medium exerts a force only at the front of the shell, then one can relate $\eta$ to the temperature of the shell and $R$, the distance the shell has traveled from the source when $\Gamma$ is a factor of 2 below its initial value. This defines a maximum value for the shell density, and therefore a minimum density for $\eta$, because when the deceleration force is spread throughout the shell, a lower pressure is required to maintain static equilibrium within the shell. Defining the dimensionless 4-velocity in the radial direction as $u_r$, the deceleration of the shell is given by \begin{equation} { {\cal M} c \over R^2 } { d u_r \over d \tau } = \Gamma p^{\prime} = \Gamma n_{shell,max}^{\prime} T \, . \end{equation} In this equation, the pressure $p^{\prime}$ is the pressure exerted on the shell by the interaction with the interstellar medium as measured in the shell rest frame. The factor of $\Gamma$ converts this proper measure of force into the radial component of the 4-acceleration. The parameter $T$ is the sum of the electron and proton temperatures in the shell in units of energy. Changing the derivative on the left-hand side of equation (3) into a derivative in $r$ and setting $dr \approx R$ and $d u \approx \Gamma$ gives a maximum density of \begin{equation} n_{shell,max}^{\prime} \approx { {\cal M} c^2 \Gamma \over 2 R^3 T } \, . \end{equation} The value of $R$ in this equation is bounded by a minimum distance $R_0$ that is set by the case in which the interstellar medium is swept up by the shell. The amount of interstellar medium per unit ster radian that must be swept up to change $\Gamma$ by a factor of 2 is $m \approx {\cal M}/\Gamma$, so \begin{equation} R_0 = \left( {3 {\cal M} \over m_p n_{ism} \Gamma } \right)^{1/3} = 3.94 \times 10^{-3} \, \pc \; {\cal M}_{27}^{1/3} n_{ism}^{-1/3} \Gamma_3^{-1/3} \, , \end{equation} where ${\cal M}_{27}$ is the shell mass per unit ster radians in units of $10^{27} \gm$, $n_{ism}$ is given in units of $\cm^{-3}$, and $\Gamma_3$ is the Lorentz factor in units of $10^{3}$. With these parameters, a shell subtending $4 \pi$ ster radians and having ${\cal M}_{27} = n_{ism} = \Gamma_3 = 1$ will carry $1.129 \times 10^{52} \ergs$ of energy, which is the characteristic value inferred from the observations at gamma-ray energies. Using equation (5) to express $R$ in units of $R_0$ in equation (4) gives \begin{equation} n_{shell,max}^{\prime} = { m_p c^2 n_{ism} \Gamma^2 \over 6 T } \, \left( {R_0 \over R} \right)^3 \, . \end{equation} The ratio of $n_{ism}$ to $n_{shell}^{\prime}$ is then \begin{equation} \eta_{min} = {n_{ism} \over n_{shell,max}^{\prime} } = { 6 T \over m_p c^2 \Gamma^2 } \, \left( {R \over R_0} \right)^3 \, . \end{equation} The value of $\eta_{min}$ is strongly dependent on the distance traveled. For deceleration over $R = R_0$ with $T = m_p c^2 \Gamma$, which is the temperature found when the interstellar medium is swept up adiabatically, one finds $\eta_{min} \approx \Gamma^{-1} = \eta_{shock}$, as expected for a shock. For lower values of $T$, one has $\eta_{min} < \eta_{shock}$ unless $R > R_0$ by a sufficiently large value. For $T \approx m_e c^2 \Gamma$, which is the case for the theory discussed below, then one has $\eta < \eta_{shock}$ for $R < 6.7 \, R_0$. Lower limits on $\eta$ for $T = m_e c^2 \Gamma$ and several different values of $\Gamma$ are given as functions of $R/R_0$ in Figures 1 and~2. \begin{figure} \figurenum{1: Limits on $\eta$} \epsscale{0.7} \plotone{f1.eps} \caption{ The variable $\eta$ is defined by equation (1) as the ratio of the interstellar medium proper density to the shell proper density. The shell deceleration distance $R$ is plotted in terms of $R_0$, which is defined in eq\hbox{.}~(5). Limits on the value of $\eta$ are plotted as solid curves for the values $\Gamma = 2 \times 10^3$, $5 \times 10^3$, and $10^4$, with the lower limits given by equation (7), and the upper limits by equation (2). The values of $\eta$ and $R$ that produce the shell thickness timescales (eq. [12]) of $0.1 \, \s$, $1 \, \s$, $10 \, \s$, and $100 \, \s$, are plotted from bottom to top as a long-dashed, dot-dashed, three-dot-dashed, and long-dashed lines, with a line for each value of $\Gamma$, which terminates on the proper boundary for that $\Gamma$. Other free parameters were set to $n_{ism} = 1 \, \cm^{-3}$, ${\cal M}_{27} = 1$.} \end{figure} \begin{figure} \figurenum{2: Limits on $\eta$} \epsscale{0.7} \plotone{f2.eps} \caption{The same as Fig\hbox{.}~1, except now $n_{ism} = 10^{5} \, \cm^{-3}$. The dotted-lines at the bottom of the figure are for a shell thickness timescale of $0.01 \, \s$, with the remaining curves the same as in Fig\hbox{.}~1.} \end{figure} The value of $\eta$ defines the light crossing time scale across the shell width. The thickness of the shell in the shell rest frame is related to the mass of the shell by \begin{equation} {{\cal M} \over R^2} = m_p n_{shell}^{\prime} l^{\prime} \, . \end{equation} Expressing $R$ in terms of $R_0$ through equation (5), one finds \begin{eqnarray} l^{\prime} &= &\left( { {\cal M} \over m_p n_{ism} } \right)^{{1\over 3}} \, \left( {\Gamma \over 3 } \right)^{{2\over 3}} \, \left({R_0 \over R}\right)^2 \, \eta \, , \\ &= &4.05 \times 10^{15} \, \cm \; {\cal M}_{27}^{{1\over 3}} \, n_{ism}^{-{1\over 3}} \, \Gamma_3^{{2\over 3}} \, \left( {R_0 \over R} \right)^2 \, \eta_{-3} \, . \end{eqnarray} where $\eta_{-3} = \eta/10^{-3}$. The thickness of the shell in the interstellar medium rest frame is $l = l^{\prime}/\Gamma$. Defining the shell's characteristic timescale $t_{shell}$ as $l = c t_{shell}$, one has \begin{eqnarray} t_{shell} &= &{\eta \over c } \left( { {\cal M} \over 9 m_p n_{ism} \Gamma} \right)^{{1\over 3}} \left({R_0 \over R}\right)^{2} \, , \\ &= &1.35 \times 10^{2} \, \s \; {\cal M}_{27}^{{1\over 3}} n_{ism}^{-{1\over 3}} \Gamma_3^{-{1\over 3}} \left( {R_0 \over R} \right)^2 \eta_{-3} \, . \end{eqnarray} For $\eta$ of order $1/\Gamma$, one requires that $R = 10 R_0$ at $n_{ism} = 1 \, \cm^{-3}$ for $t_{shell} = 1 \, \s$. Equation (12) is plotted in Figures 1 and 2 as functions of $\eta$ with respect to $R/R_0$ for several values of $\Gamma$ and $t_{shell}$. Because $t_{shell}$ must be less than or equal to the burst timescales, which are often observed to be $< 1 \, \s$, model values for $\eta$ must be $\eta \ll 1$. It is shown in \S 8 that the theory provides this limit on $\eta$. \section{Plasma Instabilities} Within the reference frame of the shell, one initially has a plane-parallel charge-neutral density profile through which a neutral and uniform plasma streams. The shell is assumed to have no initial magnetic field. The dissipation of the bulk kinetic energy of a relativistic shell to the interstellar medium must be through a plasma instability, because the densities of the shell and the interstellar medium ensure that the collision mean free path is much longer than the thickness of the shell. For a relativistic plasma, the two relevant instabilities are the two-stream instability and the electromagnetic filamentation instability (\markcite{Davidson1}Davidson 1990). For $v \ll c$, the former has a larger growth rate that the latter by a factor of $c/v$, but for $\Gamma \gg 1$, the growth rate of the latter becomes much larger than the growth rate of the former. The two-stream instability has a maximum growth rate in the relativistic regime of \begin{equation} \gamma_{2s}^{\prime} = {1 \over 2} \omega_{p, e, ism}^{\prime} \Gamma^{-{3 \over 2}} \approx {1 \over 2} \, \sqrt{ { 4 \pi e^2 \over m_e } } n_{ism}^{{1 \over 2}} \Gamma^{-1} \, , \end{equation} where $\omega_{p, e, ism}^{\prime}$ is the electron plasma frequency of the interstellar medium in the shell rest frame. Taking $\Gamma = 10^3$ and $n_{ism} = 1 \, \cm^{-3}$, one finds a growth rate for the electron two-stream instability of $\gamma_{2s, e}^{\prime} = 30 \, \s^{-1}$ in the shell frame. This growth rate corresponds to distances of travel of $10^9 \, \cm$ for the interstellar medium through the shell. The filamentation instability acts on counterstreaming plasmas by creating a magnetic pinch. Of the two plasmas, the interstellar medium and the shell, that with the smaller number density in it's own rest frame is the plasma that filaments. For gamma-ray bursts, this is the interstellar medium when $\eta \ll 1$. The growth rates of the filamentation instability is easily derived from the dielectric tensor for cold electron and ion streams in the relativistic regime and in the absence of a magnetic field. The equation must satisfy (\markcite{Davidson1}Davidson 1990) \begin{equation} 1 + \sum_{j} \left[ { \omega^{2}_{pj} \over \Gamma_j^3 c^2 k_{\perp}^2 } + { \beta_j \omega^2_{pj} \over \Gamma_j \omega^2 } \right] = 0 \, , \end{equation} where $\omega_{pj}$ is the plasma frequency of component $j$ for the density measured in the observer's rest frame, $\Gamma_j$ is this component's Lorentz factor, and $k_{\perp}$ is the wave number perpendicular to the velocity vector of the streams. If one is considering only a charge neutral stream of electrons and ions moving with Lorentz factor $\Gamma$ through a charge neutral background, and if one defines the background plasma to be the shell, then equation (14) gives a frequency of \begin{equation} \omega^{\prime \, 2} = - { \beta^2 \left( \omega_{p, e,ism}^{\prime \, 2} + \omega_{p, i, ism}^{\prime \, 2} \right) \over \Gamma \left[ 1 + c^{-2} k_{\perp}^{-2} \left( \omega_{p, e, shell}^{\prime \, 2} + \omega_{p, i, shell}^{\prime \, 2} \right) \right] } \, . \end{equation} In this equation, $\omega_{p, e,ism}^{\prime}$ and $\omega_{p, i, ism}^{\prime}$ are the electron and ion plasma frequencies of the interstellar medium, and $\omega_{p, e, shell}^{\prime}$ and $\omega_{p, i, shell}^{\prime}$ are the electron and ion plasma frequencies of the shell, all measured in the shell rest frame. The negative value of $\omega^{\prime \, 2}$ shows that the wave grows. Because $\omega_{p, e, ism}^{\prime} > \omega_{p, i, ism}^{\prime}$ and $\omega_{p, e, shell}^{\prime} > \omega_{p, i, shell}^{\prime}$, one finds that the growth rate of the filamentation of the electrons is \begin{equation} \gamma_{fe}^{\prime} = { \beta \omega_{p, e, ism}^{\prime} \over \Gamma^{1 \over 2} \sqrt{ 1 + c^{-2} k_{\perp}^{-2} \omega_{p, e, shell}^{\prime \, 2} } } \approx {1 \over 2} \, \sqrt{ { 4 \pi e^2 \over m_e } } n_{ism}^{{1 \over 2}} \, . \end{equation} The growth rate is independent of wave length for $k_{\perp} > \omega_{p,e,shell}^{\prime}/c$, and it is $\propto k_{\perp}$ otherwise. The length scale of the filament is therefore given by $x_f = k_{\perp}^{-1} = c/\omega_{p, e, shell}$. For ions alone, the growth rate, which differs from equation (16) only in the numerator, is given by \begin{equation} \gamma_{fi}^{\prime} = { \beta \omega_{p, i, ism}^{\prime} \over \Gamma^{1 \over 2} \sqrt{ 1 + c^{-2} k_{\perp}^{-2} \omega_{p, e, shell}^{\prime \, 2} } } \approx {1 \over 2} \, \sqrt{ { 4 \pi e^2 \over m_p } } n_{ism}^{{1 \over 2}} \, . \end{equation} This has the same length scale as the electron filamentation instability, but a lower growth rate. The filamentation instability growth rate is faster than the two stream instability by the factor of $\Gamma$. For $\Gamma = 10^3$, $n_{ism} = 1 \, \cm^{-3}$, one finds an electron filamentation growth rate of $\gamma_{f, e}^{\prime} = 6 \times 10^4 \, \s^{-1}$, which gives a length scale of $10^{6} \, \cm$ over which the instability occurs. For ions, the filamentation instability grows at a rate that is a factor of $\sqrt{m_e/m_p}$ slower, so that for the parameters given above, $\gamma_{f, p}^{\prime} = 1.4 \times 10^3 \, \s^{-1}$. The growth rates of all three instabilities are shown in Figures 3 and 4 as functions of $\Gamma$. Both filamentation growth rates are therefore higher than the two-stream growth rate for electrons using the parameters derived above for gamma-ray bursts. \begin{figure} \figurenum{3: Timescales} \epsscale{0.7} \plotone{f3.eps} \caption{Instability, radiative, and isotropization timescales versus Lorentz factor. Plasma timescales are given for an ISM plasma density of $n_{ism} = 1 \cm^{-3}$. The lower two solid curves are the linear growth timescales of the electron and the ion filamentation instabilities, $t_{fe}^{\prime}$ and $t_{fi}^{\prime}$. The dotted curve is the timescale for the electron two stream instability. The short dashed curve is the timescale for the electron distribution to isotropize, $t_{iso}$. The synchrotron cooling timescale $t_{sync}$ is given by the line with two dots and one dash, while the synchrotron self-Compton cooling timescale $t_{Comp}$ is given by the dash and dot line.} \end{figure} \begin{figure} \figurenum{4: Timescales} \epsscale{0.7} \plotone{f4.eps} \caption{Instability, radiative, and isotropization timescales versus Lorentz factor. As in Fig\hbox{.}~3, but for $n_{ism} = 10^{5} \cm^{-3}$.} \end{figure} \section{Filament Saturation} The filaments grow until the growth rate of the thread equals the magnetic bounce frequency of the particle beam producing the instability (\markcite{Davidson2}Davidson \etal\ 1972; \markcite{Lee}Lee \& Lampe 1973). The bounce frequency, which describes the motion of a particle across the filament through the toroidal magnetic field, is given by \begin{equation} \omega_b = \sqrt{ e B / m c \Gamma x_f } \, , \end{equation} where $x_f$ is the length scale of the filament, and $m$ is the mass of the particles comprising the filament. The size of the filaments is set by the lower limit on the wave number for which growth occurs, $k_{\perp} c = \omega_{p,e, shell}^{\prime}$. The ratio of the magnetic field energy density to the particle beam energy density in the thread is then [35] \begin{equation} { W_{B}^{\prime} \over W_{ism}^{\prime} } = { m_e n_{ism} \over m n_{shell}^{\prime} } = { m_e \over m } \, \eta \, . \end{equation} One point to note about this equation is that electron and proton components each generate magnetic fields of the same strength. For both components, one has $W_{B}^{\prime} = m_e c^2 n_{ism} \eta \Gamma^2$, which, for $\Gamma = \eta^{-1} = 10^3$, gives $B^{\prime} = 0.14 \, G$ with $n_{ism} = 1 \, \cm^{-3}$, and $B^{\prime} = 45.36 \, G$ with $n_{ism} = 10^5 \, \cm^{-3}$. This independence is a consequence of the length scale of filamentation being set by the shell electron density. The thermalization of the ions within the thread can be estimated by examining the equation of motion at saturation. A particle's equation of motion perpendicular to the thread for $u_x \ll \Gamma$ is (\markcite{Davidson2}Davidson \etal\ 1972) \begin{equation} {d^2 x \over d \tau^2 } = - \omega_b^2 x \, . \end{equation} Solving this equation of motion using a maximum spatial amplitude of $x_f$, one finds that the maximum momentum perpendicular to the filament is $u_x = {x_f \omega_b \Gamma / c }$. Replacing $x_f$ with the length scale of the filament, and replacing the bounce frequency with the filament growth rate, one finds \begin{equation} u_x = \sqrt{ W_{B}^{\prime} \over W_{ism}^{\prime} } \Gamma \, . \end{equation} If $u_x \ll 1$, then the approximate energy that goes into thermalizing the particles in the stream is approximately given by $W_{th}^{\prime}/W_{ism}^{\prime} = u_{x}^2/\Gamma$, while if $u_x \gg 1$, it is given by $W_{th}^{\prime}/W_{ism}^{\prime} = u_{x}/\Gamma$. As a result, \begin{equation} { W_{th}^{\prime} \over W_{ism}^{\prime} } = \min\left( { m_e \over m } \, \eta \Gamma, \sqrt{ { m_e \over m } \, \eta } \, \right) \, . \end{equation} A point to note is that both of the terms on the right hand side are greater than $W_{B}^{\prime}$, so that one always has $W_{th}^{\prime} > W_{B}^{\prime}$. Equations (19) and (22) are plotted in Figure 5 for $m = m_p$ and $\Gamma = 10^3$ and~$10^4$. \begin{figure} \figurenum{5: Energy fractions} \epsscale{0.7} \plotone{f5.eps} \caption{The fraction of the ISM kinetic energy, as measured in the shell rest frame, that goes into magnetic field energy (solid) and ion thermal energy (dotted) for $\Gamma = 10^3$ and $10^4$.} \end{figure} From these equations one sees that the filamentation instability cannot mediate a shock. If a shock were present, then $\eta$ would be given by equation (2). Placing this equation into equations (19) and (22) and setting $m = m_p$, one finds that ${ W_{B}^{\prime} / W_{ism}^{\prime}} \approx { m_e / m_p \Gamma } \ll 1$ and ${ W_{i}^{\prime} / W_{ism}^{\prime}} = \left( { m_e / m_p \Gamma } \right)^{1/2} \ll 1$. The energy released through the instability is small compared to the kinetic energy of the interstellar medium as measured in the shell rest frame. This contradicts the Rankine-Hugoniot equations, and so a shock never arises through the filamentation instability. \section{Electron Thermalization} Once ion filamentation is complete, electrons will attempt to come into equilibrium through the two-stream instability. This instability will drive the electrons to a distribution that is uniformly distributed in energy between the rest frame described by the ions in the filaments and the ions in the shell. The rest frame of the electron distribution must preserve both the electron charge density and the electron current, since the shell plus interstellar medium is charge-neutral. These two conditions define a rest frame for the electrons that has a Lorentz factor relative to the interstellar medium rest frame of \begin{equation} \Gamma_e = { \Gamma + \eta \over \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } } \, . \end{equation} Relative to this rest frame, the shell is moving with \begin{equation} \Gamma_s^{\prime\prime} = { \eta \Gamma + 1 \over \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } } \, , \end{equation} where the double primes are used to denote quantities measured in the electron rest frame. In the electron rest frame, the electron density is given by $n_e^{\prime\prime} = n_{ism} \Gamma_e + n_{shell}^{\prime} \Gamma_s^{\prime\prime}$, which can be written as \begin{equation} n_e^{\prime\prime} = n_{shell}^{\prime} \, \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } \, . \end{equation} The electron rest frame has nearly the same Lorentz factor as the shell rest frame as long as $\eta \ll \Gamma^{-1}$. For $1 \gg \eta \gg \Gamma^{-1}$, the Lorentz factor of the electrons is $\Gamma_e = \sqrt{ \Gamma/2 \eta }$, which is a factor of $1/\sqrt{ 2 \eta \Gamma}$ smaller than the Lorentz factor for the ions. The value $\eta = \Gamma^{-1}$ is therefore an important transition point for the character of the radiation emitted by the shell, because the Lorentz boost of the radiation is smaller than the boost associated with the shell when $\eta \gg \Gamma^{-1}$. On the other hand, because $\eta < 1$, $\Gamma_e^{\prime\prime} = \Gamma_e > \Gamma_s^{\prime\prime}$, so the two-stream instability drives the electron distribution to the energy defined by $\Gamma_e$. The electron in the shell rest frame has a Landau radius that is much larger than the filament length scale. Using the definitions for $x_f$, the gyroradius, and equation (19) for the magnetic field strength in the shell rest frame, one can write \begin{equation} {r_e \over x_f} = u_e { m_e \over m_p } \eta \, . \end{equation} The gyroradius is therefore much larger than the filament width when \begin{equation} u_e \gg {m_p \over m_e } \eta \, . \end{equation} For electrons thermalizing to the energy $\Gamma_e$, $u_e \approx \Gamma/\sqrt{ 1 + 2 \eta \Gamma }$, so that equation (27) becomes \begin{equation} \Gamma \gg \cases{ {m_p \over m_e } \eta \, , & if $\eta < \Gamma^{-1}$; \cr 2 \left( {m_p \over m_e } \right)^2 \eta^3 \, , & otherwise. \cr } \end{equation} One finds that the inequality holds for the upper term in equation (28) whenever $\Gamma > \sqrt{m_p/m_e} = 42$; below we show that $\Gamma \gtrsim 10^3$, so if equation (28) is to fail, it will be for the lower term. For $\eta \Gamma > 1$, the inequality holds as long as \begin{equation} \eta \Gamma < 2^{-{1 \over 3}} \, \left( { m_e \over m_p } \right)^{{2 \over 3}} \Gamma^{{4 \over 3}} = 52.9 \; \Gamma_3^{{4\over 3}} \, . \end{equation} When inequality (29) holds, each electron passes through many filaments in a single orbit, and the motion of the most energetic electrons will be as a single fluid, but when the inequality fails, the electrons will be confined to the local conditions within each filament, and the motion of each electron will be determined by these local conditions. For the remainder of the paper, we assume that equation (29) holds. For the electrons to flow through the magnetic fields generated in the shell, an average electric field perpendicular to the magnetic field must exist in the shell rest frame. The electric field is not uniform, since the system is charge neutral, but exists only over distances of order the width of a filament, with values that are proportional to the magnetic field strength. When the gyroradius is large, the electron effectively sees an average electric field as it completes one orbit. This electric field gives the electron guiding center a velocity of $-U_s^{\prime\prime}$ in the shell rest frame. The effective magnetic field strength in the electron rest frame is then \begin{equation} B^{\prime\prime} = {B^{\prime} \over \Gamma_s^{\prime\prime} } \, . \end{equation} Because the electron travels over many filaments, the orientation of the magnetic field changes dramatically over one gyroradius, so that the direction of the electron's velocity vector is randomized through a random-walk process rather than through the rotation over one orbit. The electron travels the filament width $x_f$ in the time $c \; \delta t^{\prime\prime}$. In this time, the angle $\theta \approx \delta t^{\prime\prime} /r_e$ is traveled, where $r_e$ is the electron gyroradius. Because the motion is a random walk, the number of time intervals $\delta t^{\prime\prime}$ required to change direction by $2 \pi$ is approximately $n = 4 \pi^2/\theta^2$. The total amount of time required to isotropize the motion of the electrons is therefore given by \begin{equation} t_{iso}^{\prime\prime} = n \Delta t^{\prime\prime} = { 4 \pi^2 r_e^2 \over c \; x_f } = { 4 \pi^2 c \Gamma_e^2 \over \omega_c^{\prime\prime \, 2} x_f } \, , \end{equation} where $\omega_c^{\prime\prime}$ is the cyclotron frequency. Using equations (19) and (30) to define the magnetic field in equation (31), one finds \begin{equation} t_{iso}^{\prime\prime} = n \Delta t = { 2 \pi^2 \Gamma_{s}^{\prime\prime} \over \omega_{p,e,ism} \eta^{3/2} } = 11.1 \, \s \; n_{ism}^{-{1 \over 2}} \eta_{-3}^{-{3\over 2}} \, { \eta \Gamma + 1 \over \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } } \, . \end{equation} For $\Gamma_{s}^{\prime\prime} \approx 1$, this timescale is long compared to the two-stream instability for electrons. \section{Burst Radiation} Two radiative processes are present within the theory: synchrotron emission and Compton scattering. The synchrotron emission will occur isotropically in the electron rest frame describe by equation (23). The observed synchrotron radiation will be boosted into the observer's reference frame by a factor of $\Gamma_e$. Compton scattering of synchrotron radiation by the synchrotron-emitting electrons boosts the radiation by another factor of $\Gamma_e^2$, because the electrons in this rest frame have a characteristic energy of $m_e c^2 \Gamma_e$. The characteristic synchrotron frequency in the shell rest frame is given by $h \nu^{\prime\prime} / m_e c^2 = 2 B^{\prime\prime}\Gamma_e^2/3B_{cr}$, where $B_{cr} = e \hbar/m_e^2 c^3$. Transforming this into the observer's reference frame and using equations (19) and (30) to remove $B^{\prime\prime}$, one finds that the characteristic synchrotron energy is \begin{eqnarray} {h \nu_s \over m_e c^2} &= &{ 2 \sqrt{ 8 \pi} \hbar e \over 3 m_e^{3/2} c^2} \eta^{{1\over2}} n_{ism}^{{1\over 2}} \, { \Gamma_e^3 \Gamma \over \Gamma_s^{\prime\prime} } \, , \\ &= &{ 2 \sqrt{ 8 \pi} \hbar e \over 3 m_e^{3/2} c^2 } \eta^{{1\over2}} n_{ism}^{{1\over 2}} \, { \left( \Gamma + \eta \right)^3 \Gamma \over \left( 1 + \eta^2 + 2 \eta \Gamma \right) \left( 1 + \eta \Gamma \right) } \, , \\ &= &2.17 \times 10^{-6} \; \eta_{-3}^{{1\over 2}} n_{ism}^{{1\over 2}} { \Gamma_3^4 \over \left( 1 + 2 \eta \Gamma \right) \left( 1 + \eta \Gamma \right) } \, , \end{eqnarray} where terms of order $\eta$ and higher were dropped in the last equation. Equation (35) places the characteristic energy in the optical band for the given characteristic energies and $\eta \Gamma < 1$. As $\eta$ increases above $\Gamma^{-1}$, the characteristic observed energy falls as $\eta^{-{3 \over 2}}$, so $\eta = \Gamma^{-1}$ defines the maximum characteristic frequency for a given value of $\Gamma$. To have a characteristic photon energy of $m_e c^2$ at $\eta = \Gamma^{-1}$ requires $\Gamma > 6.93 \times 10^4 \, n_{ism}^{-1/7}$. The characteristic energy of the Compton emission after a single scattering is a factor of $\Gamma_e^2$ larger than the synchrotron characteristic energy, so \begin{eqnarray} {h \nu_{C} \over m_e c^2} &= &{2 \sqrt{ 8 \pi} \hbar e \over 3 m_e^{3/2} c^2} \eta^{{1\over2}} n_{ism}^{{1\over 2}} { \left( \Gamma + \eta \right)^5 \Gamma \over \left( 1 + \eta^2 + 2 \eta \Gamma \right)^2 \left( 1 + \eta \Gamma \right) } \, , \\ &= &2.17 \; \eta_{-3}^{{1\over 2}} n_{ism}^{{1\over 2}} { \Gamma_3^6 \over \left( 1 + 2 \eta \Gamma \right)^2 \left( 1 + \eta \Gamma \right) } \, . \end{eqnarray} The characteristic energy of the Compton scattered radiation after a single scattering is at $m_e c^2$ when $\Gamma > 1.47 \times 10^3 \, n_{ism}^{1/11}$. A second scattering takes the photon in the observer rest frame to the GeV energy range. Further scattering does not change the photon energy, because the photon energy is of order the characteristic electron energy after the second scattering. Each of the radiative components spans a broad range of energies. The low end of the synchrotron emission is set by the cyclotron frequency, which is smaller than the characteristic synchrotron frequency by a factor of $\Gamma_e^2$. For $\eta \Gamma = 1$, the cyclotron frequency is at $\nu \approx 2.68 \times 10^8 \, \Hz \; n_{ism}^{1/2} \Gamma_3^{3/2}$, so that synchrotron emission extends down to the radio band. An important point is that the cyclotron frequency is related to the plasma frequency in the electron rest frame as $\nu_{c}^{\prime\prime} \approx \sqrt{2 } \omega_{p,e}^{\prime\prime} \eta \Gamma /2 \pi \left( 1 + 2 \eta \Gamma \right)^{1/4}$, so that they are about equal, and the cyclotron photons escape the shell to the observer. The lowest energy of the Compton scattered radiation is the cyclotron frequency photons upscattered by $\Gamma_e^2$, which means that the low end of the Compton scattered radiation equals the high end of the synchrotron emission. If most of the energy is in the Compton scattered component, and the synchrotron photon number spectrum falls faster than $\nu^{-2}$, so that most of the energy is released at the low end of the spectrum, then the low end of the Compton spectrum will be larger than the high end of the synchrotron spectrum. This implies that optical and ultraviolet emission is part of a single smooth continuum that extends through the x-ray and gamma-ray bands, and that most of the energy emitted by the burst is released in the optical and ultraviolet. The ratio of the synchrotron emission rate to the single-scattering Compton emission rate for a single electron is given by $P_{sync}^{\prime\prime}/P_{c1}^{\prime\prime} = W_{B}^{\prime\prime} / W_{synch}^{\prime\prime}$, where $W_{B}^{\prime\prime}$ and $W_{synch}^{\prime\prime}$ are the magnetic field and the synchrotron photon energy densities as measured in the electron rest frame. The synchrotron energy density is related to the single electron emission rate by $W_{sync}^{\prime\prime} = P_{synch}^{\prime\prime} n_{emis}^{\prime\prime} l^{\prime\prime}/4 \pi c$, where $n_{emis}^{\prime\prime}$ is the density of electrons with Lorentz factor $\Gamma_e$ in the electron rest frame. The synchrotron emission rate is given by $P_{sync}^{\prime\prime} = 4 \sigma_T c \Gamma_e^2 W_{B}^{\prime\prime}/3$, so the ratio becomes \begin{eqnarray} P_{sync}^{\prime\prime}/P_{c1}^{\prime\prime} &= &{ 3 \pi \left( \eta \Gamma + 1 \right) \over \sigma_T \left( \Gamma + \eta \right)^2 f_{emis} } \left( { 9 m_p \over {\cal M} n_{ism}^2 \Gamma^2 } \right)^{{1 \over 3}} \, , \\ &= &3.50 \, \Gamma_3^{-{8\over 3}} { \left( \eta \Gamma + 1 \right) \Gamma^2 \over \left( \Gamma + \eta \right)^2 } n_{ism}^{-{2 \over 3}} {\cal M}_{27}^{-{1 \over 3}} f_{emis}^{-1} \left( { R \over R_0 } \right)^2 \, . \end{eqnarray} The ratio of the synchrotron emission to the rate at which energy carried by the interstellar medium flows through the shell is given by \begin{eqnarray} { {\dot E}_{sync}^{\prime\prime} \over m_p c^3 n_{ion} \Gamma_e^2 } &= &{4 m_e \sigma_T {\cal M}^{{1\over 3}} n_{ism}^{{2 \over 3}} \Gamma^{{8\over 3}} \eta \left( 1 + \eta \Gamma \right) f_{emis} \over 3^{{5 \over 3}} m_p^{{4 \over 3}} } \left( { R_0 \over R} \right)^2 \, , \\ &= &1.956 \times 10^{-6} {\cal M}_{27}^{{1\over 3}} n_{ism}^{{2 \over 3}} \Gamma_3^{{5\over 3}} \left( \eta \Gamma \right) \left( 1 + \eta \Gamma \right) f_{emis} \left( { R_0 \over R} \right)^2 \, . \end{eqnarray} For single scattering Compton cooling, this ratio is \begin{eqnarray} { {\dot E}_{Comp}^{\prime\prime} \over m_p c^3 n_{ion} \Gamma_e^2 } &= &{2 m_e \sigma_T^2 {\cal M}^{{2\over 3}} n_{ism}^{{4 \over 3}} \Gamma^{{10\over 3}} \eta \left( \Gamma + \eta \right)^2 f_{emis}^2 \over 3^{{10 \over 3}} \pi m_p^{{5 \over 3}} } \left( { R_0 \over R} \right)^4 \, , \\ &= &2.80 \times 10^{-7} {\cal M}_{27}^{{2\over 3}} n_{ism}^{{4 \over 3}} \Gamma_3^{{13\over 3}} \left( \Gamma \eta \right) { \left( \Gamma + \eta \right)^2 \over \Gamma^2 } f_{emis}^2 \left( { R_0 \over R} \right)^4 \, . \end{eqnarray} From these equations, one sees that the energy release is very inefficient unless $n_{ism}$ or $\Gamma$ are larger than the characteristic values used in the calculations. We discuss this point in~\S 8. \section{Radiative Timescales} The cooling time scale for synchrotron emission of a relativistic electron is found by dividing the electron energy $m_e c^2 \Gamma_e$ by $P_{sync}^{\prime\prime}$, the emissivity of a single electron. In the electron rest frame, the synchrotron cooling timescale is \begin{eqnarray} t_{sync}^{\prime\prime} &= &{ 3 \over 4 \sigma_{T} c \eta n_{ism} } { \left( \eta \Gamma + 1 \right)^2 \over \Gamma^2 \left( \Gamma + \eta \right) \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } } \, , \\ &= &3.76 \times 10^{7} \, \s \; n_e^{-1} \, \Gamma_3^{-2} \, { \left( \eta \Gamma + 1 \right)^2 \over \eta \left( \Gamma + \eta \right) \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } } \, . \end{eqnarray} The synchrotron self-Compton cooling timescale is found by multiplying equations (38) and (45) together, which gives \begin{eqnarray} t_{Comp}^{\prime\prime} &= &{ 9 \pi \over 4 \sigma_{T}^2 c f_{emitt} n_{ism} } \left( { 9 m_p \over n_{ism}^2 {\cal M} } \right)^{{1 \over 3}} { \left( \eta \Gamma + 1 \right)^3 \over \eta \Gamma^{8\over 3} \left( \Gamma + \eta \right)^3 \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } } \, \left( { R \over R_0 } \right)^2 \, , \\ &= &1.32 \times 10^8 \, \s \; {\cal M}_{27}^{-{1 \over 3}} \, n_{ism}^{-{5 \over 3}} f_{emitt}^{-1} \, \Gamma_3^{-{ 14 \over 3}} \, { \left( \eta \Gamma + 1 \right)^3 \Gamma^2 \over \eta \left( \Gamma + \eta \right)^3 \sqrt{ 1 + \eta^2 + 2 \eta \Gamma } } \left( { R \over R_0 } \right)^2 \, . \end{eqnarray} These timescales are plotted in Figures 3 and 4 for $n_{ism} = 1 \, \cm^{-3}$ and $10^{5} \, \cm^{-3}$, with $\left( R/R_0 \right)^2 = m_p/m_e$, and $\eta \Gamma = 1$. One sees that for the higher densities, the Compton cooling timescale can fall below the timescale for isotropization. When this occurs, the radiative cooling will determine the shape of the electron distribution. The Compton cooling timescale is shorter than the isotropization timescale when \begin{equation} n_{ism} > 1.16 \times 10^6 \cm^{-3} \left( { R \over R_0 } \right)^{{12 \over 7}} \Gamma_3^{-{37 \over 7}} \left[ { \eta^{{1 \over 2}} \Gamma^{{7 \over 2}} \left( \eta \Gamma + 1 \right)^2 \over \left( \Gamma + \eta \right)^3 } \right]^{{6 \over 7}} \, . \end{equation} \section{Observational Consequences} The theory outlined above has within it two observational selection effects that define lower limits on the values of $n_{ism}$ and $\Gamma$. These selection effects arise because gamma-ray bursts are recognized as such through the efficient emission of gamma-rays. Because gamma-ray bursts are selected by their gamma-ray emission, one must have a value of $\eta$ that is sufficiently small to give gamma-rays through Compton scattering. Because the largest value of the photon energy occurs at $\eta = 1/\Gamma$, one can derive a lower limits on $\Gamma$ by requiring the right hand side of equation (37) be $> 1$ at this value of $\eta$: \begin{equation} \Gamma_3 > 0.831 \; n_{ism}^{-{1 \over 11}} \, . \end{equation} The weak dependence on $n_{ism}$ implies that for all gamma-ray bursts, the bulk Lorentz factor $\Gamma > 10^3$. Events may occur with smaller $\Gamma$, but these would emit in the optical and ultraviolet bands. \begin{figure} \figurenum{6: Characteristic radiation energies} \epsscale{0.7} \plotone{f6.eps} \caption{The characteristic synchrotron emission (dotted lines) and synchrotron self-Compton emission (solid lines) energies of photons emitted from the relativistic shell as measured by an observer at rest in the interstellar medium for photons emitted along the velocity vector of the shell. Values of $\Gamma = 10^3$ (upper curves) and $10^4$ (lower curves) are used. The density is $n_{ism} = 1 \, \cm^{-3}$.} \end{figure} \begin{figure} \figurenum{7: Characteristic radiation energies} \epsscale{0.7} \plotone{f7.eps} \caption{Same as Fig\hbox{.}~6, but for $n_{ism} = 10^{5} \, \cm^{-3}$.} \end{figure} From Figures 6 and 7, one sees that an upper limit on the value of $\eta$ is found from equation (37) for $\eta \Gamma \gg 1$. This limit is \begin{equation} \eta < \eta_{max} = 1.03 \times 10^{-3} \; n_{ism}^{{1 \over 5}} \, \Gamma_3^{{6 \over 5}} \, . \end{equation} These upper limits are plotted in Figures~1 and~2. An important aspect of these limits is that the timescales associated with the shell thickness are generally $< 1 \, \s$. For the higher density figure, the time scales at $R/R_0 = 0.1$ range from $0.01 \, \s$ to $1 \, \s$, which is consistent with the shortest timescales exhibited by gamma-ray bursts. Lower limits on the interstellar medium density are found by requiring that Compton scattering efficiently remove energy from the shell. Two conditions must be met: first, Compton cooling must dominate synchrotron cooling, and second, the Compton cooling rate must be comparable to the rate at which energy is lost as the shell decelerate over the distance $R$. The first of these conditions is derived from equation (39): \begin{equation} n_{ism} > 6.54 \; {\cal M}_{27}^{-{1 \over 2}} \, f_{emis}^{-{3 \over 2}} \, { \left( \eta \Gamma + 1 \right)^{{3 \over 2}} \Gamma^3 \over \left( \Gamma + \eta \right)^3 } \, \Gamma_3^{-4} \, \left( { R \over R_0 } \right)^3 \, . \end{equation} The second of these conditions is found by equating the right hand side of equation (43) to $\left(R_0/R\right)^{3} \, g$, where $g \le 1$ is a measure of efficiency. This gives \begin{equation} n_{ism} = 8.22 \times 10^{4} \, \cm^{-3} \; {\cal M}_{27}^{-{1\over 2}} \Gamma_3^{-{13\over 4}} f_{emis}^{-{3 \over 2}} \left({R \over R_0}\right)^{{3 \over 4}} \, g^{{3 \over 4}} \, . \end{equation} If $g$ is very small in equation (52), then most of the energy lost by the shell to the interstellar medium is not radiated away, making the burst dim and unobservable. \begin{figure} \figurenum{8: Limits on density} \epsscale{0.7} \plotone{f8.eps} \caption{The lower limits on $n_{ism}$ as a function of $\Gamma$.} \end{figure} The limits from equations (51) and (52) on $n_{ism}$ are plotted as functions of $\Gamma$ in Figure~8. The point to note in this figure is that the density required for high efficiency cooling is very high, of order $10^{5} \, \cm^{-3}$. This is important in explaining why all gamma-ray bursts are have a peak value of the $\nu F_{\nu}$ curve near $250 \, \keV$. The high source density provides a medium for Compton attenuation of the burst spectrum. Because the scattering medium is at rest in the galaxy rest frame, the value of $E_p$ is independent of $\Gamma$, even though the characteristic energy emitted by the shell is a strong function of $\Gamma$. As a consequence, if the medium density is low enough to keep attenuation from occurring, then the density will be too low to efficiently produce gamma-ray emission. In such a circumstance, the shell will loose energy by thermalizing ions and electrons, but the thermal energy of the electrons will not be rapidly radiated away, making these particular sources invisible. An interesting consequence of the density on the theory is that as the density increases, the timescale for radiative cooling becomes shorter than the timescale for isotropization of the electron distribution. The implications of this is that there should be a coupling of the electron distribution to the density of the interstellar medium, with the distribution falling more rapidly, and being more anisotropic, at the higher interstellar medium densities. Because the shape of the electron distribution determines the shape of the spectrum, one expects the burst spectrum to become softer for the higher interstellar medium densities, and therefore for the higher attenuation optical depths. This is the case observationally, so this theory may provide an explanation for that one characteristic of the Compton attenuation theory. The theory as constructed has two natural timescales, one from the thickness of the shell, and the second from the deceleration distance $R$. The shell thickness timescale is given by equation (12), and is of order $2.9 \, \s$ for $n_{ism} = 10^5 \cm^{-3}$ and $R/R_0 = 1$. For the same density and $R/R_0 = 10$, the timescale falls to $0.029 \, \s$. The deceleration timescale is given by \begin{equation} t_{R} = { R \over 2 c \Gamma^2} = 0.203 \, \s \; {\cal M}_{27}^{1/3} n_{ism}^{-1/3} \Gamma_3^{-7/3} \left( { R \over R_0 } \right) \, , \end{equation} where the definition of $R_0$ in equation (5) has been used. This timescale is less than the shell thickness timescale whenever \begin{equation} { R \over R_0 } > 8.74 \; \Gamma_3^{{2 \over 3}} \, \eta_{-3}^{{1 \over 3}} \, . \end{equation} Because ${ R \over R_0 }$ should be of order $m_p/m_e$, the two timescales are of the same order. When the two timescales are equal, the value is given by \begin{equation} t_{R} = t_{shell} = 1.77 \, \s \; {\cal M}_{27}^{1/3} n_{ism}^{-1/3} \Gamma_3^{-5/3} \eta_{-3}^{{1 \over 3}} \, . \end{equation} For $n_{ism} = 10^5 \, \cm^{-3}$, the timescale is $0.038 \, \s$. The timescales associated with both the shell and the deceleration distance are sufficiently short to be responsible for the shortest timescales observed in gamma-ray bursts. \section{Summary of Conclusions} To summarize the theory presented above, a baryonic shell with an ultra-relativistic bulk velocity interacts with the interstellar medium through the filamentation and the two-stream plasma instabilities. The former instability gives rise to a magnetic field with a strength that is far below the equipartition value, and the latter instability heats the electrons to energies that are relativistic, but also far below the equipartition value. Neither instability is sufficient to produce a shock. Instead, the interstellar medium passes through the shell, so that the region behind the shell is not cleared of interstellar medium. The electrons within the shell produce synchrotron radiation with a characteristic energy in the observer's rest frame that ranges from the radio to the ultraviolet. Synchrotron self-Compton emission by the electrons in the shell produces x-rays and gamma-rays in the observer's rest frame, in addition to optical emission. The optical emission is dominated by the synchrotron self-Compton component. The timescales associated with the shell thickness and the length scale over which the shell decelerates provide a lower limit on the burst durations. The burst duration itself would be determined by the complex structure of the relativistic wind, since the interstellar medium remains in place, permitting multiple shells to each produce gamma-ray emission. Two conditions must be met for the interaction between shell and interstellar medium to efficiently produce gamma-rays: first, the bulk Lorentz factor must be $> 10^3$ in order to produce radiation above $1 \keV$; second, the number density of the interstellar medium must be greater than $\approx 10^6 \cm^{-3} \, \Gamma_3^{-{13 \over 4}}$, where $\Gamma_3$ is the bulk Lorentz factor in units of $10^3$, for the thermal energy to be radiated efficiently. The lower limit on $\Gamma$ through the selection effect provides an explanation for why the value of $\Gamma$ is always sufficiently high to allow the escape of $1 \MeV$ photons from the gamma-ray burst emission region without the production of an electron-positron plasma from photon-photon pair creation and the subsequent thermalization of the radiation. The limit on density provides an explanation of why all gamma-ray burst spectra appear to be Compton attenuated. Because the limits on $\Gamma$ and $n_{ism}$ are from selection effects in observing the emission of gamma-rays, one expects there to be burst events with values of $\Gamma$ and $n_{ism}$ outside these limits. For bursts with low $\Gamma$, the bursts radiate at energies below $1 \, \keV$. Therefore, one expects a population of optical and ultraviolet transients that have no gamma-ray emission. For bursts with low density, the radiation of energy is inefficient, so that the bursts are of low intensity. These bursts may appear in burst samples through a correlation of burst intensity with the interstellar medium density inferred from the Compton attenuation model. An aspect of the theory that provides a test is the comparison of instantaneous gamma-ray, x-ray, and optical emission to the radio and optical emission. There are two aspects of the theory to test. First, one can test whether the broad band spectrum is consistent with being a synchrotron spectrum at low frequency and a Compton scattered synchrotron spectrum at higher frequency. Second, one can test the consistency of physical parameters in the theory. This last is done by comparing the Thomson and photoelectric optical depths found through a fit of the Compton attenuation model to the optical attenuation derived under the assumption that the unattenuated gamma-ray continuum extends to the optical band. Third, if the optical spectrum is sufficient to fix the cyclotron frequency by determining the low energy drop-off of the Compton spectrum, one can solve for the value of the Lorenz factor, which will provide a consistency test through the lower limit on $\Gamma$. If one can model the x-ray afterglow of a burst as the forward scattering of x-rays by dust, then one has additional information about the optical extinction that can be used in this comparison. The theory of afterglows in this theory has yet to be developed. An important difference from the shock theory of afterglows is that the region behind the shell will emit afterglow radiation in competition with the radiation from the decelerated shell. Because the evolution of the afterglow from the interstellar medium is determined by the broadening of the look-back surface and the radiative cooling of the interstellar medium, while the evolution of the shell radiation is determined by the decrease in $\Gamma$ and the evolution of the thermal structure of the shell, the theory should have two distinct afterglow components that produce a complex afterglow behavior. Numerical modeling of the plasma processes can lead to additional observational tests of the theory. In particular, the numerical modeling of the electron distribution for the regime where the electron isotropization timescale exceeds the Compton cooling timescale (Fig.~2) may provide a test through the correlation of Thomson optical depth with spectral hardness. One suspects that as $\Gamma$ increases, the synchrotron spectrum becomes softer, because an electron radiatively cools before it isotropizes, making the electron distribution one-dimensional. Because of the lower limit on $n_{ism}$ in inversely related to $\Gamma$, one expects $n_{ism}$, and therefore the Thomson optical depth, to be smaller for softer spectra. This conjecture requires numerical verification; if it is verified, then one can use the correlation of unattenuated spectral index with Thomson optical as a test of the theory. There is already some evidence of such a correlation (\markcite{Brainerd3}Brainerd \etal\ 1998, Fig. 8a). Two theoretical investigations are now warranted. The first is a study of the broad band spectrum expected for this theory need to be numerically calculated for a number of electron distributions. This study will determine what aspects of the spectrum provide tests of the theory without requiring a full understanding of the plasma processes. Such a study can be carried out through Monte Carlo simulation, and should include the effects of optical and Compton attenuation. The goal is to model the spectrum from the radio to the gamma-ray. The second is a study of the plasma processes. This will require the development of plasma codes to study the interactions of relativistic beams and the growth of instabilities to the nonlinear regime. Only such a study will verify if the analytic conclusions reached above are accurate. Only such a study will one determine the value of $\eta$ in terms of the other burst parameters. And only through such a study will one obtain model spectra that are dependent just on the bulk Lorentz factor, the density of the interstellar medium, and the mass of the relativistic shell. The plasma instability theory is capable of explaining the most important features of gamma-ray bursts. Further theoretical research is therefore justified, and should lead to strong and unambiguous observational tests of the theory.
\section{Introduction}\medskip \par There exists a strong analogy between the properties of black holes and conventional thermodynamical systems \cite{thermo}. In this analogy the entropy of the black hole is directly proportional to the surface area of its event horizon and literature refers to this quantity as the Bekenstein-Hawking entropy \cite{bekenstein}. Meanwhile, the temperature of the black hole is proportional to the surface gravity of its horizon. In spite of this established correspondence, there is still a lack of understanding of precisely what accounts for black hole entropy which is a pure geometrical quantity. In a usual statistical mechanical system the entropy is explained in terms of the degrees of freedom of its microscopic constituents. However a black hole has a limited number of such degrees of freedom, as demonstrated by the so called "No-Hair" theorems \cite{hair}. In recent literature, varied attempts have been made to derive black hole entropy on statistical mechanical principles, with varying degrees of success \cite{stringy}\cite {carlip}\cite{induced}\cite{loop}. For instance, Strominger and Vafa \cite{stringy} counted the degeneracy of soliton bound states for extremal black holes in string theory. In a very different and more geometrical approach, Carlip \cite{carlip} counted horizon edge states in a gauge theory formulation of 2+1 anti-deSitter gravity. Another approach that has been investigated involves Sakharov's theory of induced gravity \cite{sak} following a proposal by Jacobson \cite{jac}. Recent work along these lines to generate black hole entropy has been done by Frolov, Fursaev, and Zelnikov \cite{induced}. The success of a number of very diverse approaches seems to suggest that the correct explanation for black hole entropy may in some sense be universal. That is, it should depend explicitly on neither the macroscopic gravitational form nor on a hidden microscopic quantum theory \cite{wald}. Consequently, it may prove beneficial to study as wide a range of theories as feasible and in doing so look for model independent features. Such observations could potentially provide valuable insight as to the geometrical origins of black hole entropy. To this end, we examine the thermodynamic properties of black holes in generic dilaton gravity coupled to an Abelian gauge field in 1+1 dimensions. This provides a very extensive class of models which allow for black hole solutions. Even with the 2-dimensional limitation, many such models are seen to have direct physical significance. For instance, in the spherically symmetric reduction of 4-dimensional Einstein-Maxwell gravity the dilaton scalar field corresponds to radial distance. Also it has been shown that black hole solutions of constant curvature gravity in 2-dimensions ( Jackiw-Teitelboim\cite{JT}) are in fact projections of BTZ black holes described by 2+1 gravity with axial symmetry \cite{ortiz}. In recent work we have studied the classical thermodynamic properties of generic dilaton gravity via a Hamiltonian partiton function method \cite{star}\cite{shelemy}. In the present paper we calculate the thermodynamics so as to include one-loop corrections. The approach we use here is based on York's Euclidean-action method \cite{york}\cite{brown} which in turn follows from the Gibbons-Hawking path integral formalism \cite{gibbon}. This entails taking the black hole to be in a state of thermal equilibrium with evaporated radiation and then relating the periodicity of Euclidean time with the inverse thermodynamic temperature. First-order quantum corrections are introduced into the procedure via a technique applied by Frolov, Israel, and Soldukin \cite{frolov} in the study of spherically symmetric charged black holes. The basic idea is to add to the classical action a correction corresponding to the one-loop effective action obtained by integrating out matter fields coupled to the metric and non-minimally coupled to the dilaton. This one loop effective action is a suitable generalization\cite{odintsov1,hawk,kummer1,dowk} of the Polyakov action obtained from the 2-D conformal (or trace) anomaly\cite{trace}. The effect of these quantum corrections on the black hole thermodynamics is two-fold. First, it modifies the black hole geometry due to the non-vanishing one-loop effective stress-energy tensor. Secondly, the surface terms which give rise to the black hole free energy also acquire quantum corrections. As a result the formulae relevant to calculating thermodynamic quantities ( energy and entropy ) are modified as well. Our results will hopefully provide insight into the nature of such corrections for generic dilaton gravity as well as provide the template for closer examination of a myriad of specific theories. This paper is arranged as follows. In Section 2 we introduce the action for generic 2-dimensional dilaton gravity coupled to an Abelian gauge field. Here we present the most general solution to the field equations and by way of the Euclidean action approach \cite{york} we are able to describe black hole thermodynamic properties at a classical level. In Section 3 we introduce one-loop quantum corrections due to matter fields propogating on a curved background. The resulting modifications to the black hole geometry are deduced by applying the formalism of Frolov et al. \cite{frolov}. In Section 4 we calculate the quantum corrections to black hole energy and entropy. In Sections 5 and 6 we apply our results to the specific examples of charged black holes in spherically symmetric gravity and rotating BTZ black holes respectively. \footnote{Quantum corrections to the thermodynamics of the BTZ black hole and some classes of 2d charged black holes have been previously considered using different methods in \cite{odintsov2} while \cite{bytsenko} has examined quantum gravitational corrections to the entropy of the BTZ black hole.} For simplicity cases consider minimal coupling of matter fields with the dilaton however the formalism to be presented is readily extendable to more general coupling scenarios. Section 7 summarizes the paper and considers future prospects for related work. \section{Classical Theory} In two spacetime dimensions the Einstein tensor vanishes identically. Consequently, the construction of a dynamical theory of gravity with no more than two derivatives of the metric in the action requires the introduction of a scalar field, namely the dilaton. Recent works have demonstrated that the dilaton is more than a lagrange multiplier but significant in determining both the symmetries and topologies of the solution \cite{DGK}. Here we consider the most general Lorentzian action functional depending on the metric tensor $g_{\mu\nu}$, dilaton scalar field $\phi $ and Abelian gauge field $A_{\mu\nu}$ in two spacetime dimensions \cite{banks}\cite{DK}: \begin{eqnarray} W[g,\phi ,A] &=& {1\over{2G}}\int d^2x\sqrt{-g}\left[D(\phi )R(g)\right.\nonumber\\ &&\quad + \left.{1\over2}g^{\alpha\beta}\partial_{\alpha}\phi\partial_{\beta } \phi + {1\over{l^2}}V(\phi ) - {2G\over 4}Y(\phi )F^{\alpha\beta} F_{\alpha\beta}\right] \label{eq: 1} \end{eqnarray} where $G$ is the dimensionless 2-d Newton constant, $F_{\mu\nu} = \partial_{\mu}A_{\nu} - \partial_{\nu }A_{\mu }$ ,and l is a fundamental constant of dimension length. Also $D(\phi ), V(\phi ),$ and $Y(\phi )$ are arbitrary functions of the dilaton field. Variation of the action with respect to the metric, dilaton field and gauge field respectively leads to the following set of field equations. \begin{eqnarray} -2\nabla_{\alpha }\nabla_{\beta }D(\phi ) & + & \nabla_{\alpha }\phi\nabla_{\beta } \phi + g_{\alpha\beta}\left(2\Box D(\phi )-{1\over2}(\nabla\phi )^2-{1\over{l^2}} V(\phi )\right.\nonumber\\ & + & \left.{G\over 2}Y(\phi )F^{\mu\nu}F_{\mu\nu}\right)-2G Y(\phi )F^{\gamma}_{\alpha }F_{\beta\gamma} = 0 \label{eq: 2} \end{eqnarray} \begin{equation} -\Box\phi + \left[R{\delta D\over\delta\phi }+{1\over{l^2}} {\delta V\over{\delta\phi}}-{G\over 2}{\delta Y\over \delta\phi }F^ {\alpha\beta }F_{\alpha\beta }\right] = 0 \label{eq: 3} \end{equation} \begin{equation} \nabla_{\beta }(Y(\phi )F^{\alpha\beta}) = 0 \label{eq: 4} \end{equation} Directly solving Maxwell's equation \req{eq: 4} yields \begin{equation} F = {{\sqrt{-g}q}\over {Y(\phi )}} \label{eq: 5} \end{equation} where $F$ is defined implicitly by $F_{\mu r} = F\epsilon_{\mu\nu}$ and $q$ is a constant that corresponds to Abelian charge. Next we define an ``effective'' potential ${\tilde V}(\phi ,q)$ such that \begin{equation} {\tilde V}(\phi ,q) = V(\phi ) - Gl^2{q^2\over {Y(\phi )}} \label{eq: 6} \end{equation} The action and remaining field equations Eqs.[\ref{eq: 2}, \ref{eq: 3}] can be rewritten as follows: \begin{equation} W[g,\phi ,q] = {1\over {2G}}\int d^2x\sqrt{-g}\left[D(\phi )R(g) + {1\over2} (\nabla\phi )^2 + {1\over {l^2}}\tilde{V}(\phi ,q)\right] \label{eq: 7} \end{equation} \begin{equation} -2\nabla_{\alpha }\nabla_{\beta}D(\phi ) + \nabla_{\alpha}\phi\nabla_{\beta} \phi + g_{\alpha\beta}\left(2\Box D(\phi ) - {1\over2} (\nabla\phi )^2 - {1\over{l^2}}{\tilde V}(\phi , q)\right) = 0 \label{eq: 8} \end{equation} \begin{equation} -\Box\phi + R{{\delta D}\over{\delta\phi }} + {1\over l^2} {\delta{\tilde V}\over {\delta\phi }} = 0 \label{eq: 9} \end{equation} Obtaining the solutions for an action of this form has been well documented in prior works \cite{star}\cite{DK} so only a brief account will be presented here. First the action is reparameterized thereby eliminating the kinetic term (requires $D(\phi ) \neq 0$ and ${{dD}\over {d\phi }}\neq 0$ for any admissable value of $\phi $ ). \begin{equation} \overline{\phi} = D(\phi ) \label{eq: 10} \end{equation} \begin{equation} \Omega^2 = \exp{1\over2}\int{{d\phi }\over {dD/d\phi }} \label{eq: 12} \end{equation} \begin{equation} \overline{g}_{\mu\nu} = \Omega^2(\phi )g_{\mu\nu} \label{eq: 11} \end{equation} \begin{equation} {\overline{V}}(\overline{\phi} ,q) = {\tilde V}(\phi ,q)/\Omega^2(\phi ) \label{eq: 13} \end{equation} The reparameterized action is then as follows: \begin{equation} W[\overline{g} ,\overline{\phi} ,q] = {1\over {2G}}\int d^2x\sqrt{-\overline{g}} \left[\overline{\phi} R(\overline{g} ) + {1\over {l^2}}\overline{V} (\overline{\phi},q)\right] \label{eq: 13.5} \end{equation} The timelike killing vector for the resultant field equations is easily identifiable. It is found to be: \begin{equation} \overline{K}^{\mu } = {l\epsilon^{\mu\nu}\over \sqrt{-\overline{g}}}\partial_{\nu }\overline{\phi} \label{eq: 14} \end{equation} With norm given by \begin{equation} |\overline{K}|^2 = \overline{j}(\overline{\phi} ) - 2GlM \label{eq: 15} \end{equation} where \begin{equation} \overline{j}(\overline{\phi} ) = \int^{\overline{\phi} }_od\overline{\phi}\overline{V}(\overline{\phi} ,q) \label{eq: 16} \end{equation} and $M$ is a constant of integration identified as the mass observable. Next we choose a local coordinate system in which $\overline{\phi} $ and hence $\phi $ have spatial dependence only. The final solutions in these static coordinates are then obtained by exploiting the form of the killing vector. These are found to be: \begin{equation} {\overline {\phi }} = {x\over l} \label{eq: 16.25} \end{equation} \begin{equation} ds^2 = -\overline{g}(x)dt^2 + \overline{g}^{-1}(x)dx^2 \label{eq: 16.5} \end{equation} where \begin{equation} {\overline{g}}(x) = \overline{j}(\overline{\phi} ) - 2GlM \label{eq: 16.75} \end{equation} We can then re-express this solution in terms of the original parameterization as follows: \begin{equation} \phi = D^{-1}({x\over l}) \label{eq: 17} \end{equation} \begin{equation} ds^2 = -g(x)dt^2 + g^{-1}(x)\Omega^{-4}(\phi (x))dx^2 \label{eq: 18} \end{equation} where \begin{equation} g(x) = {1\over \Omega^2(\phi (x))}\left[\overline{j}({x\over l})-2GlM\right] \label{eq: 19} \end{equation} The necessary condition for a given theory to admit black hole configurations is the existence of apparent horizons. That is, spacetime curves of the form $\phi (x,t) = \phi_o$ (constant) where $\phi_o$ satisfies $g(\phi_o ; M, q) = 0$. The nature of a given black hole solution can be revealed by considering $dg / d\phi $ evaluated at these event horizons $\phi = \phi_o$. For a fixed value of mass M there may exist critical values of charge $q (M)$ so that this derivative vanishes. For such critical values the function $g(\phi ; M, q)$ may have either a local extremum or point of inflection at the horizon. If it is an extremum the norm of the killing vector does not change sign when passing through the event horizon. As $q$ is varied away from its critical value either the horizon will disappear or two event horizons (inner and outer) will appear. The latter case signifies the presence of an extremal black hole when $q$ is at its critical value. For a point of inflection the norm of the killing vector does change sign but as $q$ is varied away from its critical value one expects the formation of either one or three horizons \cite{DK}. For the subsequent (thermodynamic) analysis we consider the Euclidean sector such that $t\rightarrow it$. Hence we re-write the action \req{eq: 7} with respect to the Euclidean metric tensor: \begin{eqnarray} W_{E} &=& -{1\over 2G}\int d^2x\sqrt{g}\left[D(\phi )R(g) + {1\over 2}(\nabla\phi )^2\right.\nonumber\\ &&\quad + \left.{1\over l^2}{\tilde V}(\phi ,q)\right]-{1\over G} \oint_{outer\ boundary} dt\gamma D(\phi)\nabla_{\alpha } n^{\alpha } \label{eq: 19.5} \end{eqnarray} (Henceforth the subscript E on the Euclidean action will be implied). The second integral in \req{eq: 19.5} is the extrinsic curvature boundary term. It is included so that when second derivatives of the metric are cancelled off (via appropriate integration by parts) then the resulting total divergences on the outer boundary will be cancelled off as well \cite{gibbon}. Here we define $n_{\mu }$ as the outward unit vector normal to the outer boundary enclosing the black hole and $\gamma$ as the induced metric appropriate for evaluating the line integral. We re-write the Euclidean static metric in the following form \begin{equation} ds^2 = g(x)dt^2 + e^{-2\lambda (x)}g^{-1}(x)dx^2 \label{eq: 20} \end{equation} where $e^{\lambda(x)} = \Omega^2(\phi (x))$. For this metric it follows that \begin{equation} \sqrt{g}= e^{-\lambda (x)} \label{eq: 21} \end{equation} \begin{equation} R = -e^{\lambda (x)}\left(e^{\lambda (x)}g^{\prime}(x)\right)^{\prime} \label{eq: 22} \end{equation} where $\prime $ indicates differentiation with respect to $x$. The coordinates $t, x$ describe a disc and will be taken to range between the limits $x_+\leq x \leq L$ and $0 \leq t \leq 2\pi\beta $. Here $x = x_+$ represents the black hole horizon (ie. - $g(x_+) = 0), x = L$ is the outer boundary of the black hole (ie. - box size), and $\beta$ is the asymptotic inverse temperature. It can be shown that regularity of the solution requires the absence of a conical singularity which leads to the following condition: \begin{equation} \beta = {2e^{-\lambda (x_+)}\over g^{\prime }(x_+)} \label{eq: 23} \end{equation} Note that application of standard thermodynamics requires using the inverse temperature of the box ${\overline {\beta }}$ which is "red-shifted" from the previously defined quantity such that: \begin{equation} {\overline{\beta }} = g^{1/2}(L)\beta \label{eq: 24} \end{equation} For this metric it also follows that the extrinsic curvature (defined by the boundary term of the action) can be expressed \cite{frolov}: \begin{equation} \gamma\nabla_{\alpha }n^{\alpha} = {1\over2} e^{\lambda (L)}g^{\prime }(L) \label{eq: 25} \end{equation} Using these results Eqs.[\ref{eq: 20}-\ref{eq: 25}] we can express the Euclidean action functional (\req{eq: 19.5}), with respect to the generic static metric giving: \begin{eqnarray} W &=& {-\pi\beta\over G}\int^L_{x_+}dx\left(D^{\prime } (x)e^{\lambda (x)}g^{\prime }(x) + {e^{\lambda (x)}\over 2}g(x) (\phi^{\prime }(x))^2\right.\nonumber\\ &&\quad + \left.{e^{-\lambda (x)}{\tilde V}(x)\over l^2}\right) -{2\pi\over G}D(x_+) \label{eq: 26} \end{eqnarray} We can use this form of the action to derive thermodynamic properties of interest. These include the free energy $F = (2\pi{\overline{\beta }}) ^{-1} W $, energy $E = (2\pi )^{-1}\partial_{\overline {\beta }} W $, and entropy $S = ({\overline {\beta }}\partial_{\overline {\beta }}-1) W $. \begin{eqnarray} F &=& -{1\over 2Gg^{1/2}(L)}\int^L_{x_+}dx\left(D^{\prime }(x)e^ {\lambda (x)}g^{\prime }(x) + {e^{\lambda (x)}\over 2}g(x) (\phi^{\prime }(x))^2\right.\nonumber\\ &&\quad + \left.{e^{-\lambda (x)}{\tilde V}(x)\over l^2}\right) - {D(x_+)\over G{\overline{\beta}}} \label{eq: 27} \end{eqnarray} \begin{equation} E = -{1\over 2Gg^{1/2}(L)}\int^L_{x_+}dx\left(D^{\prime }(x)e^{\lambda (x)}g^{\prime }(x) + {e^{\lambda (x)}\over 2}g(x)(\phi^{\prime } (x))^2 + {e^{-\lambda (x)}\over l^2}{\tilde V}(x)\right) \label{eq: 28} \end{equation} \begin{equation} S = {2\pi\over G}D(x_+) \label{eq: 29} \end{equation} Since the box temperature is taken to be $T = 2\pi {\overline{\beta }}$ from Eqs.[\ref{eq: 27}-\ref{eq: 29}] we obtain the result $F = E-TS$. At the extremum of free energy (or equivalently action) $\delta F = 0$ and hence the second law of thermodynamics immediately follows. It is possible and convenient to re-express the action (\req{eq: 26}) in a form which, except for surface terms, vanishes on shell. Defining $G_{\alpha\beta }$ to be the left hand side of \req{eq: 8} and re-writing with respect to the coordinate system defined by metric \req{eq: 20} yields: \begin{eqnarray} G_{\alpha\beta } &=& -2\delta^x_{\alpha }\delta^x_{\beta } D^{\prime\prime}(x) + 2\Gamma^x_{\alpha\beta }D^{\prime }(x) + g_{\alpha\beta }\left[2e^{\lambda (x)}\left(e^{\lambda (x)}g(x) D^{\prime }(x)\right)^{\prime }\right.\nonumber\\ &&\quad - \left.{1\over2} g(x)e^{2\lambda (x)}(\phi^{\prime } (x))^2-{1\over l^2}{\tilde V}(x)\right]=0 \label{eq: 30} \end{eqnarray} In the case where both tensor indices represent time coordinate denoted by 0 (and note that $\Gamma^x_{00} = -{1\over2} e^{2\lambda}gg^{\prime })$ we get: \begin{eqnarray} G^0_0 &=& -e^{\lambda(x)}\left[-e^{\lambda (x)}g^{\prime }(x)D^{\prime }(x) - 2g(x)\left(e^{\lambda (x)}D^{\prime }(x)\right)^{\prime }\right.\nonumber\\ &&\quad +\left. {1\over2} g(x)e^{\lambda (x)}(\phi^{\prime }(x))^2 + {e^{-\lambda (x)}\over l^2}{\tilde V}(x)\right] \label{eq: 31} \end{eqnarray} Using this result to substitute for the second and third terms in the integrand of \req{eq: 26} : \begin{equation} W = -{\pi\beta\over G}\int^L_{x_+}dx\left[-e^{-\lambda (x)} G^0_0+ 2\left(g(x)e^{\lambda (x)}D^{\prime }(x)\right)^{\prime } \right] - {2\pi\over G}D(x_+) \label{eq: 32} \end{equation} Since the second term in the integrand is a total derivative and $g(x_+) = 0$ it follows that: \begin{equation} W = {\pi\beta\over G}\int^L_{x_+}e^{-\lambda (x)}G^0_0dx - {2\pi\beta\over G}e^{\lambda (L)}g(L)D^{\prime }(L) - {2\pi\over G}D(x_+) \label{eq: 33} \end{equation} Reconsider the energy $E = (2\pi )^{-1}\partial_{{\overline{\beta }}} W $. Since thermodynamic quantities are presumed to be calculated for equilibrium configurations (i.e. on shell) here we can set $G^0_0 = 0$ giving an energy which reduces to an outer boundary surface term: \begin{equation} E = -{1\over G}e^{\lambda (L)}g^{{1\over2} }(L)D^{\prime }(L) \label{eq: 34} \end{equation} This expression is typically divergent as $ L\rightarrow\infty $. (This follows from the divergence of the Euclidean action as the outer boundary goes to infinity). To resolve this dilemma we compare the energy of \req{eq: 34} with that of a carefully selected background geometry \cite{hawk2}. The background metric will be taken here to represent the asymptotic geometry of the black hole. Hence we define $ g_0 = lim_{L\rightarrow\infty} g(L) $ and the "subtracted energy" is given by: \begin{equation} E_{sub} = {1\over G}e^{\lambda(L)}D^{\prime}(L)\left[g^{{1\over2}}_0-g^{{1\over2}}(L)\right] \label{eq: 34.5} \end{equation} We can justify this choice of background by noting the agreement between this result with that attained for a Hamiltonian partition function approach in a prior study \cite{star}. \section{Quantum Corrected Black Hole Geometry} In the path integral approach to black hole thermodynamics the matter fields can be integrated out yielding an effective action which depends only on fields in the classical action. Hence one-loop quantum effects can be taken into account by adding a quantum counterpart $\hbar\Gamma $ to the classical gravitational action $W_{CL}$ (\req{eq: 19.5}) such that (assuming no matter coupling to the Abelian gauge field): \begin{equation} W\left[g,\phi,q\right] = W_{CL}\left[g,\phi,q\right] + \hbar\Gamma \left[g,\phi\right] \label{eq: 35} \end{equation} Variation of this complete action yields the quantum corrected field equations which may be solved perturbatively. Variation of the action with respect to the metric gives us: \begin{equation} G_{\alpha\beta }(g,\phi,q) + \hbar T_{\alpha\beta }(g,\phi) + O(\hbar^2) = 0 \label{eq: 35.5} \end{equation} where $G_{\alpha\beta } = {\delta W_{CL}\over \delta g^{\alpha\beta }}$ is again given by the left hand side of \req{eq: 8} and $T_{\alpha\beta } = {\delta\Gamma\over \delta g^{\alpha\beta }}$. The general form of the one loop effective action is \cite{odintsov1,kummer1}: \begin{equation} \Gamma = {1\over 12}\int d^2x\sqrt{g}\left[aR{1\over \Box}R +b(\phi)(\nabla\phi)^2{1\over \Box}R +c(\phi) R - \ln(\mu^2)b(\phi) (\nabla\phi)^2\right] \label{eq: 36} \end{equation} The first term is the usual trace anomaly that arises for minimally coupled scalars, while the next two terms are contributions to the anomaly from the non-minimal coupling to the dilaton. The last term contains an arbitrary scale factor, $\mu$, and comes from the conformally invariant part of the effective action\footnote{We are grateful to S. Odintsov for pointing out the necessity for including this term.}. It can be obtained using a Schwinger-DeWitt type expansion\cite{odintsov2}. We will show that the scale factor $\mu$ does not affect the final thermodynamic quantities. $a$ is a constant (we will set $a=1$ for simplicity), while $b(\phi)$ and $c(\phi)$ are determined by the specific form of the coupling between the matter fields and the dilaton. For example, in dimensionally reduced spherically symmetric gravity, $b$ and $c$ are constants\cite{hawk,dowk,odintsov1,kummer1}. Here we will treat them as arbitrary local functions of the dilaton. It is important for the following thermodynamic analysis to put the non-local expression \req{eq: 36} for $\Gamma $ in local form. We do this by introducing a pair of scalar fields $\psi$ and $N$, and writing: \begin{eqnarray} \Gamma &=& {1\over 12}\int d^2x\sqrt{g}\left[(\psi + N) R + (\nabla N) \cdot(\nabla\psi )+b(\nabla\phi)^2(\psi-\ln(\mu^2))+c\phi R \right] \nonumber\\ &&\quad + {1\over 6} \oint_{outer\ boundary}dt\gamma(\psi +N+c(\phi))\nabla_{\alpha } n^{\alpha } \label{eq: 37} \end{eqnarray} where an extrinsic curvature surface term has been added in analogy to the classical case. It is straightforward to show that variation of \req{eq: 37} yields the following field equations for the scalars: \begin{equation} \psi = {1\over \Box } R \label{eq: psi1} \end{equation} and \begin{equation} N={1\over \Box } ( R+b(\nabla \phi )^2 ) \label{eq: N1} \end{equation} Substituting these equations back into \req{eq: 37} yields precisely \req{eq: 36}. Note that in the 2-dimensional minimally coupled case ($b=0$), N reduces to $ \psi $ and only a single scalar field need be introduced. Before proceeding we show it is possible to solve explicitly for $\psi (x)$ at the classical level (as is appropriate for this analysis). This is achieved by conformally mapping the coordinate space described by the static Euclidean metric \req{eq: 20} to a flat disc of radius $z_o$ and curvature $ R=\Box\psi $. This disc may be expressed \begin{equation} ds^2 = e^{-\psi (z)}(z^2d\theta^2 + dz^2) \label{eq: 38} \end{equation} where the disc coordinates $\theta $ and $z$ are taken to range between $0 \leq \theta \leq 2\pi $ and $0 \leq z \leq z_0$. Substituting the Euclidean static metric \req{eq: 20} into the left hand side gives \begin{equation} g(x)dt^2 + g^{-1}(x)e^{-2\lambda (x)}dx^2 = {\overline {e}}^{\psi(z)}(z^2d\theta^2 + dz^2) \label{eq: 39} \end{equation} where $t = \beta\theta $ and $x_+ \leq x \leq L$. The following relations follow directly from \req{eq: 39}: \begin{eqnarray} g(x)dt^2 &=&e^{-\psi }z^2d\theta^2\nonumber\\&=& {e^{-\psi }z^2\over \beta^2}dt^2 \label{eq: 40} \end{eqnarray} \begin{equation} g^{-1}(x)e^{-2\lambda }dx^2 = e^{-\psi }dz^2 \label{eq: 41} \end{equation} Using \req{eq: 40} to solve for $z$ and \req{eq: 41} to solve for $dz$ gives us: \begin{equation} z = \beta\sqrt{g}e^{+\psi /2} \label{eq: 42} \end{equation} \begin{eqnarray} dz&=&{e^{-\lambda}e^{+\psi/2}\over\sqrt{g}}dx\nonumber\\ &=& {e^{-\lambda}z\over \beta g}dx \label{eq: 43} \end{eqnarray} Dividing \req{eq: 43} by \req{eq: 42} and integrating for given boundary conditions yields: \begin{equation} ln \left({z_o\over z}\right) = {1\over \beta }\int^L_x{dx\over g(x)} e^{-\lambda (x)} \label{eq: 44} \end{equation} Using equation \req{eq: 42} to re-write the left-hand side of \req{eq: 44} as a function of $x$ and solving for $\psi (x)$ : \begin{equation} \psi (x) = \psi (L) - {2\over \beta} \int^L_x dx{e^{-\lambda (x)}\over g(x)} + ln g(L) - lng(x) \label{eq: 45} \end{equation} To find an explicit expression for $\psi (L)$ consider the calculation of the proper time evaluated for a closed path on the boundary of the disc at $x = L(z = z_o)$: \begin{equation} \oint^{2\pi\beta }_{t=0}\sqrt{g(L)}dt = \oint^{2\pi }_{\theta = 0} z_o e^{-\psi (L)/2}d\theta \label{eq: 46} \end{equation} Integrating and solving for $\psi (L)$ : \begin{equation} \psi (L) = -2ln\left({\beta\over z_o}\right)-lng(L) \label{eq: 47} \end{equation} Substituting \req{eq: 47} into \req{eq: 45}: \begin{equation} \psi (x) = -lng(x) -{2\over \beta }\int^L_xdx{e^{-\lambda (x)}\over g(x)}-2ln\left({\beta\over z_o}\right) \label{eq: 48} \end{equation} To solve for N we repeat the prior analysis except here we map to a flat disc described in the form: \begin{equation} ds^2=e^{-N(z)+{1\over \Box}\left[b(\nabla \phi)^2\right]}(z^2d\theta^2+dz^2) \label{eq:48.01} \end{equation} This results in the following: \begin{equation} N(x)=-lng(x)-{2\over\beta}\int^L_xdx{e^{-\lambda(x)}\over g(x)} -2ln({\beta\over z_0})+{1\over\Box}\left[b(\phi)\left(\nabla\phi (x)\right)^2 \right] \label{eq:48.02} \end{equation} Because of the non-local form of the last term this is not a satisfactory result so we integrate $ \Box N = R -b(\nabla\phi)^2 $ giving \begin{eqnarray} N(x)&=& N(L)-lng(x)+lng(L)\nonumber\\ &&-\int^L_xd{\tilde x}{e^{-\lambda({\tilde x} )}\over g({\tilde x})}\left[C-\int^L_{\tilde x}d{\overline {x}}\,b\, e^{\lambda({\overline{x}})}(\phi^{\prime}({\overline{x}}))^2 g({\overline{x}})\right] \label{eq: 48.0999} \end{eqnarray} where $N(L)$ and $C$ are arbitrary constants of integration. The constant $N(L)$ does not affect the thermodynamic quantities in the subsequent analysis, so without loss of generality we set $N(L)=\psi(L)$. The remaining constant must in principle be determined by experiment. However, we adopt the ansatz that $N(x)$ should reduce to $\psi(x)$ when $b=0$, and that the geometry should uniquely determine both $\psi$ and $N$. With these conditions N reduces to: \begin{equation} N(x)=\psi(x)+\int^L_xd{\tilde x}{e^{-\lambda({\tilde x})}\over g({\tilde x})}\int^L_{\tilde x}d{\overline {x}}\,b\,e^{\lambda({\overline{x} })}(\phi^{\prime}({\overline{x}}))^2g({\overline{x}}) \label{eq: 48.1} \end{equation} If we signify $g_{CL}$ as the classical metric and $g = g_{CL} + \delta g$ as the one-loop quantum corrected metric it can be shown (by perturbative expansion) that the following form of the field equation \req{eq: 35.5} is valid to first order: \begin{equation} G_{\alpha\beta }(g) + \hbar T_{\alpha\beta }(g_{CL}) = 0 \label{eq: 49} \end{equation} Where $G_{\alpha\beta }$ is given by the left hand side of \req{eq: 8} and $T_{\alpha\beta }$ can be obtained from the variation of \req{eq: 37}. We find: \begin{eqnarray} T_{\alpha\beta }&=&{G\over 3} \left[ \nabla_{\alpha }\nabla_{\beta }(\psi +N) - {1\over 2}(\nabla_{\alpha}N\nabla_{\beta}\psi+ \nabla_{\alpha }\psi\nabla_{\beta }N)\right.\nonumber\\ &&\quad\quad-g_{\alpha\beta } (R+\Box N-{1\over2} (\nabla N )\cdot(\nabla\psi)) \nonumber\\ &&\quad\quad -b(\psi-\ln(\mu^2))(\nabla_{\alpha}\phi\nabla_{\beta}\phi-{1\over2} g_{\alpha\beta}(\nabla\phi)^2)\nonumber\\ &&\left.\quad\quad-(g_{\alpha\beta}\Box c(\phi)-\nabla_{\alpha} \nabla_{\beta}c(\phi)) \right] \label{eq: 50} \end{eqnarray} An explicit expression for $ T_{\alpha\beta} $ in terms of the metric is then obtained by substituting for $ \psi $ and N via \req{eq: 48} and \req{eq: 48.1} respectively. It should be noted that the resulting equation can be equivalently obtained by direct functional differentiation of the action in its non-local form (\req{eq: 36}) \cite{torre}. We again take the dilaton as representing the spatial coordinate so that the geometric corrections are manifested in the metric. Solving the field \req{eq: 49} yields an explicit form of the quantum corrected metric. In analogy to the formalism presented by Frolov et al. \cite{frolov} we adapt the classical static metric (Eqs.[\ref{eq: 18}-\ref{eq: 19}]) to the quantum corrected case as follows: \begin{equation} ds^2 = g(x)e^{2w(x)}dt^2 + g^{-1}(x)\Omega^{-4}(\phi (x)) dx^2 \label{eq: 51} \end{equation} \begin{equation} g(x) = {1\over \Omega^2(\phi (x))}\left[\overline{j} ({x\over l}) -2GlM - 2Glm(x)\right] \label{eq: 52} \end{equation} Here $m(x)$ is the first order quantum correction to the classical mass $M$ and we have introduced a metric function $w(x)$ which vanishes in the classical limit (and where functions $\overline{j} $ and $\Omega^2$ are as defined by Eqs.[\ref{eq: 12},\ref{eq: 13},\ref{eq: 16}]). We now solve $G_{\alpha\beta } = -\hbar T_{\alpha\beta }$ by first finding expressions for quantum quantities $m(x)$ and $w(x)$ in terms of components of the tensor $T_{\alpha\beta }$ . Using \req{eq: 8} for $ G_{\alpha\beta} $ gives us: \begin{equation} -2\nabla_{\alpha }\nabla_{\beta }D + \nabla_{\alpha }\phi\nabla_{\beta }\phi + g_{\alpha\beta }\left[2\Box D -{1\over2}(\nabla\phi )^2-{1\over l^2}{\tilde V}\right] = -\hbar T_{\alpha\beta } \label{eq: 53} \end{equation} Using the fact that the solution only depends on $x$ and the definition of covariant derivative: \begin{equation} -2\delta^x_{\alpha }\delta^x_{\beta }D^{\prime \x}+2\Gamma^x_{\alpha\beta }D^{\prime } + \delta^x_{\alpha }\delta^x_{\beta }(\phi^{\prime })^2 + g_{\alpha\beta } \left[2\Box D-{g^{xx}\over 2}(\phi^{\prime })^2-{1\over l^2} {\tilde V}\right] = -\hbar T_{\alpha\beta } \label{eq: 54} \end{equation} The off diagonal components (i.e. $\alpha=x$, $\beta=t$) of the above equation vanish identically. For the case in which both indices $\alpha ,\beta $ represent the time coordinate: \begin{equation} -g^{xx}g^{\prime }_{tt}D^{\prime }+g_{tt}\left[2\Box D-{g^{xx}\over 2}(\phi^{\prime })^2-{1\over l^2}{\tilde V}\right] = -\hbar T_{tt} \label{eq: 55} \end{equation} Now we re-express the left hand side with respect to the metric defined by Eqs.[\ref{eq: 51}-\ref{eq: 52}]. First note that by using $D = {x\over l}$ (\req{eq: 17}) we can evaluate $\Box D$ to give \begin{equation} \Box D = {\Omega^2\over l}(\Omega^2gw^{\prime } + \overline{j}^{\prime }-2Glm^{\prime }) \label{eq: 56} \end{equation} and so: \begin{eqnarray} e^{2w}\left[-{2\Omega^4g^2w^{\prime }\over l}-{\Omega^2g\over l}(\overline{j}-2Glm)^{\prime }+{g^2\Omega^2\over l}(\Omega^2)^{\prime }\right.&&\nonumber\\ \left.+{2g\Omega^2\over l}(g\Omega^2w^{\prime }+\overline{j}^{\prime }-2Glm^{\prime })- {\Omega^4g^2\over 2}(\phi^{\prime })^2-{g\over l^2} {\tilde V}\right] &=& -\hbar T_{tt} \label{eq: 57} \end{eqnarray} From Eqs.[\ref{eq: 13},\ref{eq: 16}] \begin{equation} \overline{j}^{\prime} = {1\over l}{\tilde V\over \Omega^2} \label{eq: 58} \end{equation} and from Eqs.[\ref{eq: 12},\ref{eq: 17}] : \begin{equation} (\Omega^2)^{\prime } = {l\Omega^2\over 2}(\phi^{\prime })^2 \label{eq: 59} \end{equation} Using these 2 results in \req{eq: 57} and solving for $m^{\prime }$ (and using $T_{tt} = g_{tt}T^t_t)$ : \begin{equation} m^{\prime } = {\hbar\over 2G\Omega^2}T^t_t \label{eq: 60} \end{equation} Now for the case in which both tensor indices in \req{eq: 54} represent the spatial coordinate: \begin{equation} -2D^{\prime \x }+g^{xx}g^{\prime }_{xx}D^{\prime }+(\phi^{\prime })^2+g_{xx}\left[2\Box D-{g^{xx}\over 2}(\phi^{\prime })^2-{1\over l^2} {\tilde V}\right]=-\hbar T_{xx} \label{eq: 61} \end{equation} Using the metric (Eqs.[\ref{eq: 51}-\ref{eq: 52}]) along with \req{eq: 56} : \begin{eqnarray} &- &{1\over lg\Omega^2}(\overline{j} -2GlM)^{\prime }-{(\Omega^2)^{\prime }\over l\Omega^2}+{(\phi^{\prime })^2\over 2}\nonumber\\ &&\quad+{2\over lg\Omega^2} (g\Omega^2w^{\prime }+\overline{j}^{\prime }-2Glm^{\prime }) - {{\tilde V}\over l^2g\Omega^4} = -\hbar T_{xx} \label{eq: 62} \end{eqnarray} Using Eqs.[\ref{eq: 58},\ref{eq: 59}] and solving for $w^{\prime }$ : \begin{equation} w^{\prime } = {l\over 2}\left({2G\over g\Omega^2}m^{\prime }-\hbar T_{xx}\right) \label{eq: 63} \end{equation} Substitute for $m^{\prime }$ via \req{eq: 60} and use $T_{xx} = g_{xx}T^x_x$ : \begin{equation} w^{\prime } = {l\hbar\over 2g\Omega^4}(T^t_t - T^x_x) \label{eq: 64} \end{equation} \req{eq: 60} and \req{eq: 64} provide the first order quantum corrections to the geometry. Note that consistency of the perturbative expansion requires $T_{\alpha\beta }$ in the above expressions to be evaluated on the classical solution. Next we explicitly evaluate the tensor components $T^t_t$ and $T^x_x$. In terms of the classical static metric as expressed by \req{eq: 20} the non-vanishing terms are given by ( after some simplification ): \begin{eqnarray} T^t_t &=& {G\over 3} \left[{1\over2} g^{\prime}e^{2\lambda}(\psi^{\prime} +N^{\prime})+{1\over2} g e^{2\lambda}N^{\prime}\psi^{\prime}+2 e^{\lambda} (e^{\lambda}g^{\prime})^{\prime}\right.\nonumber\\ &&\quad\left.-b g e^{2\lambda}(\phi^{\prime})^2 (1-{1\over2}(\psi-\ln(\mu^2))-e^{2 \lambda}( gc^{\prime\prime} +gc^{\prime}\lambda^{\prime}+{1\over2} g^{\prime}c^{\prime}) \right] \label{eq: 65} \end{eqnarray} \begin{eqnarray} T^x_x&=&{G\over 3}\left[g e^{2\lambda}(\psi+N)^{\prime\prime}+({1\over2} g^{\prime}+g\lambda^{\prime})e^{2\lambda}(\psi +N)^{\prime}-{1\over2} g e^{2\lambda}N^{\prime}\psi^{\prime}+2 e^{\lambda} (e^{\lambda}g^{\prime})^{\prime}\right.\nonumber\\ &&\quad\left.- b g e^{2\lambda}(\phi^{\prime})^2 (1+{1\over2}(\psi-\ln(\mu^2))-{1\over2} e^{2\lambda}g^{\prime}c^{\prime} \right] \label{eq: 66} \end{eqnarray} Substituting for $\psi $ (\req{eq: 48}) and $N$ (\req{eq: 48.1}) and further simplification gives: \begin{eqnarray} T^t_t &=&{G e^{2\lambda}\over 6 g}\left[ 4g e^{-\lambda}(e^{\lambda}g^ {\prime})^{\prime}-(g^{\prime})^2+{4 \over \beta^2}e^{-2\lambda} \right.\nonumber\\ &&\quad\quad- 2 b g^2(\phi^{\prime})^2\left( 1+{1\over2} ln g +{1 \over\beta}\int^L_x dx{e^ {-\lambda}\over g}+ln({\beta\over z_0\mu})\right)\nonumber\\ &&\quad\quad\left. -2{e^{-2\lambda}\over\beta} \int^L_x dx be^{\lambda}g(\phi^{\prime})^2 -2 g \left(gc^ {\prime\prime}+gc^{\prime}\lambda^{\prime}+{1\over2} g^{\prime}c^{\prime }\right)\right] \label{eq: 67} \end{eqnarray} \begin{eqnarray} T^x_x &=&{G e^{2\lambda}\over 6 g}\left[(g^{\prime})^2-4{e^{-2\lambda} \over\beta^2}\right.\nonumber\\ &&\quad \quad- 2 bg^2(\phi^{\prime})^2\left( -{1\over2} ln g-{1\over\beta} \int^L_x dx{ e^{-\lambda}\over g}-ln({\beta\over z_0\mu})\right)\nonumber\\ &&\quad\quad\left. +2{e^{-2 \lambda}\over\beta} \int^L_x dx be^{\lambda}g(\phi^{\prime})^2 -g g^{\prime}c^{\prime}\right] \label{eq: 68} \end{eqnarray} Substituting these results into \req{eq: 60} and \req{eq: 64} gives us the desired explicit expressions for the first order quantum corrected mass $M(x) = M_{CL}+m(x)$ and metric function $w(x)$. Integrating and using $\Omega^2 = e^{\lambda }$ leads to the results: \begin{eqnarray} M(x)&=&M_{CL}+{\hbar\over 6}\int^xdxe^{\lambda }\left[ 2e^{-\lambda }(e^{\lambda }g^{\prime })^{\prime }-{(g^{\prime })^2 \over 2g}+{2 e^{-2\lambda }\over \beta^2g}\right.\nonumber\\ &&\quad\quad- bg(\phi^{\prime})^2 \left(1+{1\over2} ln g+{1\over\beta}\int^L_x dy{e^{-\lambda(y)}\over g(y)}+ln( {\beta\over z_0\mu})\right)\nonumber\\ &&\quad\quad\left.-{e^{-2\lambda}\over\beta g}\int^L_x dy b e^ {\lambda(y)}g(\phi^{\prime}(y))^2 -\left(gc^{\prime\prime} +gc^{\prime}\lambda^{\prime}+{1\over2} g^{\prime}c^{\prime}\right)\right] \label{eq: 69} \end{eqnarray} \begin{eqnarray} w(x) &=& -{l\hbar G\over 6}\int^L_xdx{1\over g}\left[2 e^{-\lambda }(e^{\lambda }g^{\prime })^{\prime }-{(g^{\prime })^2\over g}+ {4 e^{-2\lambda }\over \beta^2g}\right.\nonumber\\ &&\quad \quad- 2 bg(\phi^{\prime})^2\left({1\over2}+{1\over2} ln g +{1\over\beta}\int^L_x dy {e^{-\lambda(y)}\over g(y)}+ln({\beta\over z_0\mu})\right) \nonumber\\ &&\quad\quad\left. -2{e^{-2\lambda}\over\beta g}\int^L_x dy be^{\lambda(y)}g(y)(\phi^{\prime}(y)) ^2 -g\left(c^{\prime\prime}+c^{\prime}\lambda^{\prime} \right)\right] \label{eq: 70} \end{eqnarray} Here we have imposed the condition $w(L) = 0$ and have absorbed the lower limit of \req{eq: 69} into the constant $M_{CL}$. Also of importance (particularly for the evaluation of the quantum corrected entropy) is evaluation of the first order quantum shift in the horizon and hence in the horizon value of the dilaton field. To this purpose we define $\Delta\phi_+ = \phi_+ - \phi_{+CL} $ where $\phi_+ $ and $\phi_{+CL} $ are the quantum corrected and classical horizon values of the dilaton field respectively. Because the norm of the killing vector (\req{eq: 15}) must vanish at the horizon it follows that the above fields must satisfy: \begin{equation} \overline{j} (\phi_{+})-2M_{CL}Gl-2m(\phi_+)Gl = 0 \label{eq: 71} \end{equation} \begin{equation} \overline{j} (\phi_{+CL})-2M_{CL}Gl = 0 \label{eq: 72} \end{equation} Expanding $\overline{j} (\phi_+)$ about $\phi_{+CL}$ (to first order) and using \req{eq: 58} to evaluate the derivative of $ {\overline{j}} $ gives: \begin{equation} \overline{j} (\phi_+) = \overline{j} (\phi_{+CL})+{1\over l}\left.{\tilde {V}(\phi_+)\over \Omega^2(\phi_+)}{1\over \phi^{\prime }} \right|_{\phi_{+CL}} \Delta\phi_+ \label{eq: 73} \end{equation} Substituting for the $\overline{j} $'s via Eqs.[\ref{eq: 71},\ref{eq: 72}] and solving for $\Delta\phi_+$ gives to first order (note $m\sim\hbar $): \begin{equation} \Delta\phi_+ = \left.{2Gl^2m(\phi )\Omega^2(\phi )\phi^{\prime }\over \tilde {V}(\phi )}\right|_{\phi_{+CL}} \label{eq: 74} \end{equation} \section{Quantum Corrections to Black Hole Thermodynamics} Here we calculate the thermodynamical quantities $E = (2\pi )^{- 1}\partial_{{\overline {\beta }}}W$ and $S = ({\overline {\beta }} \partial_{{\overline {\beta }}}-1)W$ for the action functional \req{eq: 35} which describes the one loop quantum corrected black hole configuration: \begin{eqnarray} W &=& W_{CL}\left[g\right] + \hbar\Gamma\left[g\right]\nonumber\\ &=& W_{CL}\left[g_{CL}\right] + \hbar {\delta W_{CL} \over \delta g}\vert_{{g_{CL}}}\delta g + \hbar\Gamma \left[g_{CL}\right] + O\left[\hbar^2\right] \label{eq: 75} \end{eqnarray} Recall \req{eq: 33} for the classical action $ W_{CL}[g_{CL}]$. This included a term with an integrand proportional to $G^0_0$ and an inner and outer surface term. It is possible and convenient to derive an analogous expression for the quantum effective action $\Gamma $. Rewriting \req{eq: 37} for $\Gamma $ in terms of the static classical metric \req{eq: 20}, using \req{eq: 25} to evaluate the extrinsic curvature boundary term and integrating by parts leads to: \begin{eqnarray} \Gamma &=&{\pi\beta\over 6}\int^L_{x_{+}}dx\left[e^{\lambda}g^{\prime} \left(\psi +N+c\right)^{\prime}+e^{\lambda}g N^{\prime}\psi^{\prime}+b e^ {\lambda}g(\phi^{\prime})^2(\psi-\ln(\mu^2))\right]\nonumber\\ &&\quad+{\pi\over 3}\left( \psi(x_+)+N(x_+)+c(x_+)\right) \label{eq: 76} \end{eqnarray} Now recall \req{eq: 65} for $ T^0_0 = T^t_t $ . Re-writing this result ( making use of definitions of $ \Box\psi $ and $ \Box N $ and rearranging ) : \begin{eqnarray} T^0_0 &=& {Ge^{\lambda }\over 6}\left[e^{\lambda}g^{\prime}\left(\psi+N+c\right) ^{\prime}+e^{\lambda}g N^{\prime}\psi^{\prime}+b e^{\lambda}g(\phi ^{\prime})^2(\psi-\ln(\mu^2))\right.\nonumber\\ &&\quad\left.+ 4(e^{\lambda}g^{\prime})^{\prime}+2\left(e^{\lambda}g (\psi-N)^{\prime}\right)^{\prime}-2 (e^{\lambda}gc^{\prime})^{\prime} \right] \label{eq: 77} \end{eqnarray} Using this result to substitute for the integrand in \req{eq: 76} yields : \begin{eqnarray} \Gamma &=&{\pi\beta\over G}\int^L_{x_{+}}dx\left[ e^{-\lambda}T^0_0-{2 G\over 3}(e^{\lambda}g^{\prime})^{\prime} -{G\over 3}\left(e^{\lambda}g(\psi-N)^{\prime}\right)^{\prime} +{G\over 3}(e^{\lambda}gc^{\prime})^{\prime}\right]\nonumber\\ &&\quad\quad\quad+ {\pi\over 3}\left(\psi (x_+)+N(x_+)+c(x_+)\right) \label{eq: 78} \end{eqnarray} Since three of the four terms in the integrand are total derivatives we get : \begin{eqnarray} \Gamma ={\pi\beta\over G}\int^L_{x_{+}}dx e^{-\lambda }T^0_0 &-&{2\pi\beta\over 3}e^{\lambda (L)}g^{\prime }(L)\nonumber\\ &-&{\pi\beta\over 3}e^{\lambda(L)}g(L)\left(\psi^{\prime}(L)- N^{\prime}(L)-c^{\prime}(L)\right)\nonumber\\ &+& {\pi\over 3}\left(\psi(x_+)+N(x_+)+c\phi(x_+)\right) \label{eq: 79} \end{eqnarray} where we have used $g^{\prime }(x_+)e^{\lambda (x_+)}=2/\beta $ and discarded the irrelevant constant term which results. When we combine this result for $\Gamma $ with the first order quantum corrected form for $W_{CL}$ into \req{eq: 75} we obtain an integral with integrand proportional to $G^0_0(g)+\hbar T^0_0(g_{CL})$ along with boundary terms. Since the integrand vanishes on shell according to the field equation (\req{eq: 49}) we are left with only surface contributions to $W_{on\ shell}$. These are found to be \begin{eqnarray} W_{on\ shell} &=& -2\pi \left[{\beta\over G}e^{\lambda (L)}g(L)D^{\prime }(\phi (L))+{D\over G}(\phi (x_+))+{\hbar\over 3}\beta e^{\lambda (L)}g^{\prime }_{CL}(L)\right.\nonumber\\ &&\quad\quad + {\hbar\over 6}\beta e^{\lambda(L)}g_{CL} (L)\left(\psi^{\prime}(L)-N^{\prime}(L)-c^{\prime}(L)\right)\nonumber\\ &&\quad\quad\left.-{\hbar\over 6}\left(\psi(x_+)+N(x_+)+c_{CL}(x_+)\right)\right] \label{eq: 80} \end{eqnarray} where the surface contributions from $W_{CL}$ are obtained by generalizing \req{eq: 33}. Note that $g(x)$ and $\phi (x_+)$ in the first two terms refer to the quantum corrected solutions whereas the remaining terms are defined with respect to classical geometry. Evaluation of thermodynamic quantities is then straightforward giving: \begin{eqnarray} E = - {e^{\lambda (L)}\over G}g^{{1\over2} }(L)D^{\prime }(\phi (L)) &-&{\hbar\over 3}e^{\lambda (L)}g^{-{1\over2} }_{CL}(L)g^{\prime }_{CL}(L) \nonumber\\ &+&{\hbar\over 6}e^{\lambda(L)}g^{{1\over2}}_{CL}(L)c^{\prime}(L) \label{eq: 81} \end{eqnarray} \begin{equation} E_{sub}=E(g;g_{CL})-E(g_0;g_{0CL}) \label{eq: 81.5} \end{equation} \begin{eqnarray} S &=& {2\pi\over G}D(\phi (x_+))-{\hbar\over 6}2\pi \left[2\psi(x_+)+c_{CL}(x_+)\right.\nonumber\\ \quad\quad&&+\left.\int^L_{x_{+}}dx{e^{-\lambda(x)}\over g_{CL}(x)} \int^L_{x}d{\overline{x}}be^{\lambda({\overline{x}})} \phi^{\prime}({\overline{x}})^2g({\overline{x}})\right] \label{eq: 82} \end{eqnarray} Where $ g_{0} $ and $ g_{0CL} $ represent the background geometry and are the metric fields evaluated at $ x=L\rightarrow\infty $. Note that the left-most terms in the expressions for energy and entropy have classical forms but have implied quantum corrections due to geometry . On the other hand, the remaining terms all vanish in the classical ( $ \hbar \rightarrow 0 $ ) limit. \section{Quantum Corrections in Spherically Symmetric Reduced Gravity Theory} Next we want to use the preceding formalism to examine a specific theory. Here we consider the form of action obtained from the spherically symmetric reduction of 4-dimensional Einstein-Maxwell gravity to a 2- dimensional dilaton model \cite{SSG}. We will specifically examine the minimal case $ b=c=0 $ so that $ N=\psi $. This case in particular was studied by Frolov et al. \cite{frolov} and we find our results to be in agreement. We proceed by considering an effective action of the following form (Note that we neglect writing the extrinsic curvature term for sake of brevity but its inclusion is implied.): \begin{equation} W_{CL} = -{1\over 2Gl^2}\int d^2x\sqrt{g}\left[{r^2\over 2}R(g)+ (\nabla r)^2+\left(1-Q^2/r^2\right)\right] \label{eq: 83} \end{equation} Where $Gl^2 = G^{(4)}$ is the square of the 3+1 dimensional Planck length and where the ``effective'' charge $Q$ has dimensions of length. Comparison with the form of the classical action (\req{eq: 13.5}) leads to the following identifications: \begin{equation} \phi = {\sqrt{2}r\over l} \label{eq: 84} \end{equation} \begin{equation} D(\phi ) = D(r) = {r^2\over 2l^2} \label{eq: 85} \end{equation} \begin{equation} {\tilde V}(\phi ,q) = {\tilde V}(r, Q) = 1-{Q^2\over r^2} \label{eq: 86} \end{equation} The classical solution Eqs.[\ref{eq: 17}-\ref{eq: 19}] can then be expressed: \begin{equation} x = {r^2\over 2l} \label{eq: 87} \end{equation} \begin{equation} \Omega^2(r) = e^{\lambda (r)} = {r\over l} \label{eq: 88} \end{equation} \begin{equation} {\overline {j}}({x\over l}) = \overline{j} (r) = {r\over l}\left(1+Q^2/r^2\right) \label{eq: 89} \end{equation} \begin{equation} g(r) = 1-{2Gl^2M\over r}+{Q^2\over r^2} \label{eq: 90} \end{equation} Note that the constant of integration in the evaluation of the integral defined in \req{eq: 12} for $\Omega^2$ is selected to be $-2\ln\sqrt{2}$. This yields a metric function $g(r)$ which goes to 1 as $r\rightarrow\infty $, which is the correct asymptotic behaviour of the metric in spherically symmetric gravity. For subsequent calculations it will often be convenient to express the metric function $g(r)$ in the following form (Here we consider solutions only for which two real, distinct horizons exist, i.e. $(l^2GM)^2>Q^2$ .) \begin{equation} g(r) = {1\over r^2}(r-r_+)(r-r_-) \label{eq: 91} \end{equation} where $r_{\pm }$ represents the outer(+) and inner (-) horizons given by: \begin{equation} r_{\pm } = l^2GM\pm\sqrt{(l^2GM)^2-Q^2} \label{eq: 92} \end{equation} Before proceeding to evaluate the quantum corrected quantities, we consider the classical energy and entropy. The classical energy (\req{eq: 34}) in terms of $r$ for this theory becomes: \begin{eqnarray} E &=& -{1\over G}e^{{\lambda }(r=L)}g^{{1\over2} }(r=L) {dD(r)\over dr}\vert_{r=L}{dr\over dx}\nonumber\\ &=& -{L\over Gl^2}\sqrt{1-{2Gl^2M\over L}+{Q^2\over L^2}} \label{eq: 93} \end{eqnarray} Above and for the remainder of this section $L$ is taken to be the value of $r$ at the outer boundary. Clearly this energy is divergent as $L\rightarrow\infty $. Hence we apply the standard subtraction procedure as defined by \req{eq: 34.5} to give: \begin{equation} E_{sub} = {L\over Gl^2}\left(1-\sqrt{1-{2Gl^2M\over L}+{Q^2\over L^2}}\right) \label{eq: 94} \end{equation} Taking the asymptotic ($L \rightarrow \infty $) limit yields the expected result $\lim_{L \rightarrow \infty }(E_{sub}) = M$. The classical entropy (\req{eq: 29}) is: \begin{equation} S = {\pi r^2_+\over Gl^2} \label{eq: 95} \end{equation} Next we calculate the quantum corrected quantities in spherically symmetric theory. Recall we consider the minimal case $ b=c=0 $ so that $ N=\psi $ . To avoid confusion classical-specific quantities will be labelled with the subscript ``CL''. Re-writing \req{eq: 69} for quantum corrected mass $M(x)$ in terms of $r$ and substituting \req{eq: 91} for the classical metric gives us: \begin{eqnarray} M(r) &=& M_{CL}+{\hbar\over 6}\int^{r}dr\left[2{d^2g_{CL} \over dr^2}-{({dg_{CL}\over dr})^2\over 2g_{CL}}+ {2\over \beta^2_{CL}g_{CL}}\right]\nonumber\\ &=& M_{CL}+{\hbar\over 6}\int^rdr\left[{10\over r^4}(r- r_{+CL})(r-r_{-CL})-{6\over r^3}(2r-r_{+CL}-r_{-CL})\right. \nonumber\\ &&\quad+ {3\over r^2} - {(r-r_{-CL})\over 2r^2(r-r_{+CL})}-{(r-r_{+CL}) \over 2r^2(r-r_{-CL})}\nonumber\\ &&\quad\left.+{2r^2\over (r-r_{+CL})(r-r_{- CL})\beta^2_{CL}}\right] \label{eq: 96} \end{eqnarray} Integrating and using \req{eq: 23} \begin{equation} \beta_{CL} =\left. {2e^{-\lambda (r)}\over \left({dg_{CL}\over dr} \right)\left({dr\over dx}\right)}\right|_{r=r_{+CL}} = {2r^2_{+CL}\over (r_{+CL}-r_{-CL})} \label{eq: 97} \end{equation} gives us the following \begin{equation} M(r) = M_{CL} + {\hbar\over 6}\left[A\ln(r-r_{-CL})+B\ln(r) + C(r)\right] \label{eq : 98} \end{equation} where: \begin{equation} A = {(r_{+CL}-r_{-CL})^2(r_{+CL} +r_{-CL}) (r^2_{+CL} +r^2_{-CL})\over 2r^4_{+CL} r^2_{-CL}} \label{eq: 99} \end{equation} \begin{equation} B = - {(r_{+CL}-r_{-CL})^2(r_{+CL}+r_{-CL})\over 2r^2_{+CL} r^2_{-CL}} \label{eq: 100} \end{equation} \begin{equation} C(r) = {2r\over \beta^2_{CL}}+{(r_{+CL}-r_{-CL})^2\over 2rr_{+CL}r_{-CL}}+{2(r_{+CL}+r_{-CL})\over r^2}-{10r_{+CL} r_{-CL}\over 3r^3} \label{eq: 101} \end{equation} Note that $A+B = 4M_{CL}Gl^2/\beta^2_{CL}$. Consider the quantum corrected mass for some special cases. For $M(r = L)$ for large $L$ then $\ln(L-r_{-CL})\sim\ln(L)$ and using the prior property for $A$ and $B$ gives: \begin{equation} M(L)\sim M_{CL}+{\hbar\over 3\beta^2_{CL}}[L+2M_{CL}Gl^2\ln(L)] \label{eq: 102} \end{equation} Also consider the case of an uncharged black hole. The classical metric function becomes $g_{CL} = (1-r_{+CL}/r)$ where $r_{+CL} = 2M_{CL}Gl^2$ and $\beta_{CL} = 2r_{+CL}$. An analogous calculation to that presented above then gives for the uncharged case: \begin{equation} M(r) = M_{CL}+{\hbar\over 12}\left[{r\over r_{+CL}^2}+ {7r_{+CL}\over 2r^2}-{1\over r}+{\ln(r)\over r_{+CL}}\right] \label{eq: 103} \end{equation} For the case of an extremal black hole $r_{-CL}\rightarrow r_{+CL}$ and $\beta_{CL}\rightarrow\infty $. Consequently $A$ and $B$ vanish and $ C(r) $ reduces to: \begin{equation} C(r) = 4r_{+CL}/r^2-10r_{+CL}^2/3r^3 \end{equation} Next consider the metric function $w(x)$. Re-writing \req{eq: 70} in terms of $r$ and then substituting \req{eq: 91} for the classical metric, using \req{eq: 97} for the inverse asymptotic temperature and finally integrating gives the result \begin{equation} w(r) = {\hbar Gl^2\over 6}\left(F(L)-F(r)\right) \label{eq: 104} \end{equation} where : \begin{eqnarray} F(r) &=& -{\left[3r_{+CL}^2+2r_{+CL}r_{-CL}+3r^2_{- CL}\right]\over r_{+CL}^2r^2_{-CL}}\ln(r)\nonumber\\ &&\quad + {\left[3r_{+CL}^4+2r_{+CL}^3r_{-CL}+2r_{+CL}^2 r^2_{-CL}+2r_{+CL}r^3_{-CL}-r^4_{-CL}\right]\over r^4_{+CL}r^2_{-CL}}ln(r-r_{-CL})\nonumber\\ &&\quad + {4\over r^2}+{4(r_{+CL}+r_{-CL})\over rr_{+CL}r_{-CL}} -{(r_{+CL}^4-r^4_{-CL})\over r_{+CL}^4r_{-CL}(r-r_{-CL})} \label{eq: 105} \end{eqnarray} As for the quantum corrected mass we consider some special cases. For an uncharged black hole the function $F(r)$ takes the simpler form: \begin{equation} F(r) = {3\over 2r^2}+{2\over rr_{+CL}}-{1\over r_{+CL}^2}\ln(r) \label{eq: 106} \end{equation} If $L$ is large, we can write: \begin{equation} e^{2w(r)}\sim \left({r\over L}\right)^{{\hbar Gl^2\over 3r_{+CL}^2}} \exp\left(-{\hbar Gl^2\over 3}\left({3\over 2r^2}+{2\over rr_{+CL}}\right)\right) \label{eq: 107} \end{equation} In the extremal black hole limit the function $F(r)$ reduces to: \begin{equation} F(r) = {8\over r_{+CL}^2}\ln ({r-r_{+CL}\over r} )+{4\over r^2}+{8\over rr_{+ CL}} \label{eq: 108} \end{equation} Consequently at the extremal black hole horizon $F(r_+)\rightarrow -\infty $ so that $e^{2w(r_+)}\rightarrow\infty $. Next we examine the quantum corrected energy. Revising \req{eq: 81} for reduced spherically symmetric gravity: \begin{eqnarray} E &=& -{L\over Gl^2}\sqrt{1-{2Gl^2M(L)\over L}+{Q^2\over L^2}}\nonumber\\ &&\quad-{\hbar\over 3} {1\over L^2}\left(2Gl^2M_{CL}-{2Q^2\over L} \right) {1\over \sqrt{1-{2Gl^2M_{CL}\over L}+{Q^2\over L^2}}} \label{eq: 109} \end{eqnarray} Where $M(r=L)$ is given by \req{eq: 103}. Consider the case of large box size $L$. Clearly the second part of the expression is small relative to the first. So we consider the first term only and substitute for the quantum corrected mass (for large $r = L$) by way of \req{eq: 102}: \begin{equation} E\sim -{L\over Gl^2}\sqrt{1-{2Gl^2M_{CL}\over L}-{2\hbar Gl^2\over 3\beta^2_{CL}} -{4\hbar(Gl^2)^2M_{CL}\over 3\beta^2_{CL}L}\ln(L)+{Q^2\over L^2}} \label{eq: 110} \end{equation} As in the calculation of classical energy we again apply the standard subtraction procedure of comparing the divergent quantity with that of a background defined by the metric $g_0 = \lim_{L\rightarrow\infty}g(r=L)$. Since $g(L)$ for large $L$ is the quantity inside the square root sign in \req{eq: 110} it follows that $g_0 = 1 - {2\hbar Gl^2\over 3\beta^2_{CL}}$ and since $E[g_0] = -{Lg^{{1\over2} }_0\over Gl^2}$ the subtracted energy is given by: \begin{eqnarray} E_{sub} &\sim& {L\over Gl^2}\left[\sqrt{1-{2\hbar Gl^2\over 3\beta^2_{CL}}}\right.\nonumber\\ &&\left.- \sqrt{1-{2Gl^2M_{CL}\over L}-{2\hbar Gl^2\over 3\beta^2_{CL}} -{4\hbar (Gl^2)^2M_{CL}\ln(L)\over 3\beta^2_{CL}L} +{Q^2\over L^2}}\right] \label{eq: 111} \end{eqnarray} The approach we use here is to first fix $L$ and expand the square roots with respect to the perturbative factor $\hbar .$ Then we take $L$ to be large and expand with respect to ${1\over L}.$ Then eliminating all $O\left({1\over L^2}\right)$ terms inside the square brackets leaves: \begin{equation} E_{sub} \sim M_{CL} + {\hbar Gl^2M_{CL}\over 3\beta^2_{CL}}\left(2\ln (L) + 1\right) \label{eq: 112} \end{equation} Note that the first order quantum correction to the energy can be attributed to temperature effects since $\beta_{CL}$ represents the asymptotic inverse temperature. Next consider the quantum corrected value of the horizon radius $r_+$. For this purpose we define $\Delta r_+ = r_+-r_{+CL}$ and as previously defined $m(r) = M(r)-M_{CL}$. The quantum corrected metric must vanish at $r = r_+$ and this relation can be expressed as follows: \begin{equation} 0 = g(r_+) = g_{CL}(r_+) - {2Gl^2m(r_+)\over r_+} \label{eq: 113} \end{equation} Expanding to first order about $r_+ = r_{+CL}$ and using $g_{CL}(r_{+CL})=0$ and expressing $(dg_{CL}/dr)_{r=r_{+CL}}$ in terms of $\beta_{CL}$ (\req{eq: 97}) gives: \begin{equation} \Delta r_{+}= {\beta_{CL}G l^2 m(r_{+CL})\over r_{+CL} } \label{eq: 114} \end{equation} Note that $m(r_{+CL})$ contains a factor of $\hbar $( see \req{eq: 69}). From this result we can calculate the quantum correction to the horizon area which is proportional to $r^2_+ = (r_{+CL} + \Delta r_+)^2$ and hence to first order: \begin{equation} r^2_+ = r_{+CL}^2 + 2\beta_{CL}Gl^2m(r_{+CL}) \label{eq: 115} \end{equation} Finally in this section we evaluate the quantum correction to the entropy. For this theory the entropy (\req{eq: 82}) is: \begin{equation} S = {\pi r^2_+\over Gl^2} - \hbar {2\pi\over 3}\psi (r_{+CL}) \label{eq: 116} \end{equation} Making use of the preceding result for $r^2_+$ (\req{eq: 115}) we can write: \begin{equation} S = S_{CL} + 2\pi\beta_{CL}m(r_{+CL}) - \hbar {2\pi\over 3}\psi (r_{+CL}) \label{eq: 117} \end{equation} Revising \req{eq: 48} for this theory gives us: \begin{equation} \psi (r_{+CL}) = -\ln g_{CL}(r_{+CL})-{2\over \beta_{CL}}\int^L_{r_+} {dr\over g_{CL}(r)}-2\ln\left({\beta_{CL}\over z_0}\right) \label{eq: 118} \end{equation} Using \req{eq: 91} for $g_{CL}$, \req{eq: 97} for $\beta_{CL}$, integrating the middle term, and simplification yields: \begin{eqnarray} \psi (r_{+CL}) &=& {r^2_{-CL}\over r_{+CL}^2}\ln\left( {L-r_{-CL}\over r_{+CL}-r_{-CL}}\right)-\ln\left( {L-r_{+CL}\over r_{+CL}-r_{-CL}}\right)\nonumber\\ &&\quad - {(r_{+CL}-r_{-CL})\over r_{+CL}^2}(L-r_{+CL}) -2\ln\left({r_{+CL}\over z_0}\right) \label{eq: 119} \end{eqnarray} The third term in $\psi (r_{+CL})$ can be interpreted\cite{frolov} as the contribution to the entropy of a two-dimensional hot gas in a box size $L-r_{+CL}$ and temperature $(2\pi\beta_{CL})^{-1} = (r_{+CL}-r_{-CL})/4\pi r_{+CL}^2$. Hence we subtract off this contribution to obtain the quantum corrected black hole entropy: \begin{eqnarray} S &=& S_{CL}+2\pi\beta_{CL}m(r_{+CL})-{\hbar 2\pi\over 3}{r^2_{-CL} \over r_{+CL}^2}\ln\left({L-r_{-CL}\over r_{+CL}-r_{-CL}}\right)\nonumber \\ & &\qquad+\hbar {2\pi\over 3}\ln \left({L-r_{+CL}\over r_{+CL}-r_{-CL}}\right) + \hbar {4\pi\over 3}\ln \left( {r_{+CL}\over z_0}\right) \label{eq: 120} \end{eqnarray} We next consider some special cases. In the case of large box size $L$ the entropy reduces to: \begin{equation} S \sim S_{CL}+2\pi\beta_{CL}m(r_{+CL})+{2\pi\over 3}\left (1-{r^2_{-CL}\over r_{+CL}^2}\right)\ln\left( {L\over r_{+CL}-r_{-CL}}\right) + {4\pi\over 3}\ln \left({r_{+CL}\over z_0}\right) \label{eq: 121} \end{equation} For an uncharged black hole then \begin{equation} S = S_{CL} + 2\pi\beta_{CL}m(r_{+CL}) + {2\pi\over 3} \ln \left({Lr_{+CL}\over z^2_0}\right) \label{eq: 122} \end{equation} where $m(r_{+CL})$ can be evaluated using \req{eq: 103}. Finally, in the extremal black hole limit \begin{equation} S = S_{CL} + 2\pi\beta_{CL}m(r_{+CL}) + {2\pi\over 3}\ln \left({r_{+CL}^2\over z^2_0}\right) \label{eq: 123} \end{equation} where $m(r_{+CL}) = \hbar/9r_{+CL}$ in the extremal case however $\beta_{CL}\rightarrow\infty $ so the entropy is divergent in this limit.\\ \section{Quantum Corrections in Jackiw-Teitelboim Theory} In this section we examine the Achucarro-Ortiz black hole \cite{ortiz}, which is a solution to the field equations for Jackiw-Teitelboim gravity\cite{JT}. This theory can be obtained by imposing axial symmetry in 2+1 dimensional gravity, so that the Achucarro- Ortiz black hole corresponds to the projection of the BTZ axially symmetric black hole \cite{BTZ} into 1+1-dimensional spacetime. The Jackiw-Teitelboim field equations can be derived from an effective action of the form \cite{ortiz} \begin{equation} W_{CL} = -\int d^2x\sqrt{g}\Lambda^{{1\over2} }\left[rR(g) + \Lambda r -{J^2\over 2r^3}\right] \label{eq: 124} \end{equation} where $\Lambda $ is the cosmological constant (dimension length$^{-2}$) and $J$ is an ``effective charge'' (dimension length) which describes the angular momentum of the $BTZ$ black hole. Note that there is no kinetic term in this action so it is of the form (\req{eq: 13.5}) without the need for a field reparametrization. This leads to the following identification (provided we set 2G = 1 and $l = \Lambda^{-{1\over2} }$): \begin{equation} \overline{\phi} = \Lambda^{{1\over2} }r \label{eq: 125} \end{equation} \begin{equation} D(\overline{\phi} ) = D(r) = \Lambda^{{1\over2} }r \label{eq: 126} \end{equation} \begin{equation} {\tilde V}(\overline{\phi} ,q) = {\tilde V}(r, J) = \Lambda^{{1\over2} } (r - {J^2\over 2\Lambda r^3}) \label{eq: 127} \end{equation} The classical solution Eqs.[\ref{eq: 16.25}--\ref{eq: 16.75}] can then be expressed: \begin{equation} x = r \label{eq: 128} \end{equation} \begin{equation} \overline{j} (\overline{\phi} ) = \overline{j} (r) = {\Lambda r^2\over 2}+{J^2\over 4r^2} \label{eq: 129} \end{equation} \begin{equation} g(r) = {\Lambda r^2\over 2}-{M\over \Lambda^{{1\over2} }}+{J^2\over 4r^2} \label{eq: 130} \end{equation} Here we consider only solutions for which two real, distinct horizons exist (i.e. $M^2 > \Lambda^2J^2/2$) so it will prove convenient to express the metric \req{eq: 130} in the following form \begin{equation} g(r) = {\Lambda\over 2}{(r^2-r^2_+)(r^2-r^2_-)\over r^2} \label{eq: 131} \end{equation} where $r_{+}$ and $r_-$ represent the inner and outer horizons, respectively, given by: \begin{equation} r^2_{\pm } = {1\over \Lambda^{3/2}}\left[M\pm \sqrt{M^2-\Lambda^2J^2/2}\right] \label{eq: 132} \end{equation} Note that because the action is already in reparameterized form we set $\Omega^2 = 1$ (or equivalently $\lambda = 0$) in the previously derived results. The classical, subtracted energy for this theory, as described by \req{eq: 34.5}, is given by: \begin{equation} E_{sub}=2^{{1\over2}}\Lambda L\left(1-\sqrt{1-{2M\over\Lambda^{{3\over 2} }L^2}+{J^2\over 2\Lambda L^4}}\right) \label{eq: 133.5} \end{equation} whle the classical entropy (\req{eq: 29}) is: \begin{equation} S = 4\pi\Lambda^{{1\over2} }r_+ \label{eq: 134} \end{equation} Next we calculate the quantum corrected quantities in Jackiw- Teitelboim theory with minimal coupling ($ b=c=0 $) as for SSG in the prior section. Henceforth, purely classical quantities will be labelled with the subscript ``CL.'' \req{eq: 69} for the quantum corrected mass $M(x=r)$ gives us (substituting for classical metric \req{eq: 131}): \begin{eqnarray} M(r) &=& M_{CL}+{\Lambda\hbar\over 6}\int^rdr \left[{4\over r^2}(r_{+CL}^2+r^2_{-CL})\right.\nonumber \\ & &\quad\left.+{5\over r^4} (r^2-r^2_{+CL})(r^2-r^2_{-CL}) -{(2r^2-r_{+CL}^2-r^2_{-CL})\over (r^2- r_{+CL}^2)}\right.\nonumber\\ &&\quad - \left.{(2r^2-r_{+CL}^2-r^2_{-CL})\over (r^2-r^2_{-CL})}+{4\Lambda^{-2}r^2\over \beta^2_{CL}(r^2- r_{+CL}^2)(r^2-r^2_{-CL})}\right] \label{eq: 135} \end{eqnarray} Integrating and using via \req{eq: 23} \begin{equation} \beta_{CL} = {2\over \left({dg_{CL} \over dr }\right)}\vert_{r=r_{+CL}} = {2\over \Lambda r_{+CL} \left(1-r^2_{-CL}/r_{+CL}^2\right)} \label{eq: 136} \end{equation} gives us the following result \begin{equation} M(r) = M_{CL}+{\hbar\Lambda\over G}\left[A\left(\ln(r-r_{-CL})-\ln (r+r_{-CL})\right)+B(r)\right] \label{eq: 137} \end{equation} where: \begin{equation} A = {r_{+CL}^6-3r_{+CL}^4r_{-CL}^2+3r_{+CL}^2r_{-CL}^4-r_{-CL}^6 \over 2r_{+CL}^2r_{-CL}(r_{+CL}^2-r_{-CL}^2)} \label{eq: 138} \end{equation} \begin{equation} B(r)=r+{r_{+CL}^2\over r}+{r_{-CL}^2\over r}-{5\over 3}{r_{+CL }^2r_{-CL}^2\over r^3} \label{eq: 139} \end{equation} Next consider the quantum corrected mass for some special cases. For$M(r=L)$ and large $L$ then coefficient of $A\sim 0$ and we are left with: \begin{equation} M(L) \sim M_{CL}+{\hbar\over 6}\Lambda L \left[ 1+{r_{+CL}^2 +r_{-CL}^2\over L^2}\right] \label{eq: 140} \end{equation} Also we consider the ``chargeless'' case. The classical metric function becomes $g_{CL}={\Lambda\over 2}(r^2-r_{+CL}^2)$ where $r_{+CL}^2 = 2M\Lambda^{-3/2}$ and $\beta_{CL} = 2/\Lambda r_{+CL}$. Repeating the above calculation yields: \begin{equation} M(r) = M_{CL} + {\hbar\over 6}\Lambda r \label{eq: 141} \end{equation} For the extremal black hole case $r_{-CL}\rightarrow r_{+CL}$ and $\beta_{CL}\rightarrow\infty $ . The coefficient of $A$ vanishes and $B(r)$ reduces to: \begin{equation} B(r)=r+2{r_{+CL}^2\over r}-{5\over 3}{r_{+CL}^4\over r^3} \label{eq: 141.5} \end{equation} Next consider the metric function $w(x=r)$. Applying \req{eq: 70} for this theory and using \req{eq: 131} for the classical metric and \req{eq: 136} for the inverse asymptotic temperature gives: \begin{equation} w(r) = {\hbar\over 6\Lambda^{{1\over2} }}(F(L)-F(r)) \label{eq: 142} \end{equation} where: \begin{equation} F(r)={4\over r}-{r(r_{+CL}^2-r_{-CL}^2)\over r_{+CL}^2(r^2- r_{-CL}^2)}+ln({r-r_{-CL}\over r+r_{-CL}})\left[ { 3r_{+CL}^4 -2r_{+CL}^2r_{-CL}^2-r_{-CL}^4 \over 2r_{-CL}r_{+CL}^2( r_{+CL}^2-r_{-CL}^2)}\right] \label{eq: 143} \end{equation} For the uncharged case we find using the revised metric function discussed above that $w(r) = 0$ for all allowable $r$. Meanwhile for the extremal limit $F(r)$ reduces to $4/r$ except at the horizon where the right most term in \req{eq: 143} becomes a divergent quantity. Hence in this limit $F(r_+)\rightarrow-\infty $ and as in the preceding section the factor $e^{2w(r_+)}$ is divergent. Next we consider the quantum corrected energy. Revising \req{eq: 81} for Jackiw-Teitelboim theory: \begin{equation} E = -2^{{1\over2}}\Lambda L\left[\sqrt{1-{2M(L)\over\Lambda^{{3\over 2}}L^2 }+{J^2\over 2\Lambda L^4}}+{\hbar\over 3\Lambda^{{1\over2}}L}{ (1 - {J^2\over2\Lambda L^4}) \over \sqrt{1-{2M_{CL}\over \Lambda^{3\over 2 }L^2}+ {J^2\over 2\Lambda L^4}}}\right] \label{eq: 144} \end{equation} Applying the usual background subtraction procedure (\req{eq: 81.5}) then gives: \begin{eqnarray} E_{sub}&=&2^{{1\over2}}\Lambda L\left[1-\sqrt{1-{2M(L)\over\Lambda^ {{3\over 2}}L^2}+{J^2\over 2\Lambda L^4}}\right.\nonumber\\ \quad\quad&&\left.+{\hbar\over 3\Lambda^{{1\over2}}L}\left(1- {(1-{J^2\over 2\Lambda L^4})\over\sqrt{1-{2M_{CL}\over\Lambda}^{{3\over 2}}L^2}+{J^2\over 2\Lambda L^4}}\right)\right] \label{eq: 144.5} \end{eqnarray} In the case of large box size $L$ the second part vanishes relative to the first. Consequently, for large $L$ the primary contribution to the quantum shift in energy is a result of the shift in mass as described by \req{eq: 140}. So to first order in $\hbar $ and to zero'th order in ${1\over L}$ we find: \begin{equation} E_{sub}\sim (E_{sub})_{CL} + {2^{{1\over2} }\hbar\Lambda^{{1\over2} }\over 6} \label{eq: 145} \end{equation} Following the procedure for calculating the quantum correction to the horizon radius $\Delta r_+ = r_+-r_{+CL}$ which was introduced in the previous section we find \begin{equation} \Delta r_+ = {\beta_{CL}m(r_{+CL})\over 2\Lambda^{{1\over2} }} \label{eq: 146} \end{equation} where $m(r) = M(r)-M_{CL}$ is given by \req{eq: 137}. Furthermore the first order quantum correction to the horizon area can be obtained from: \begin{equation} r^2_+ = r_{+CL}^2 + {\beta_{CL}r_{+CL}M(r_{+CL})\over \Lambda^{{1\over2} }} \label{eq: 147} \end{equation} Finally, in this section we determine the quantum correction to entropy. For Jackiw-Teitelboim theory the dilaton generic entropy (\req{eq: 82}) becomes: \begin{equation} S = 4\pi\Lambda^{{1\over2} }r_+ - {\hbar 2\pi\over 3}\psi (r_{+CL}) \label{eq: 148} \end{equation} Making use of the preceding result $\Delta r_+$ (\req{eq: 146}): \begin{equation} S = S_{CL}+2\pi\beta_{CL}m(r_{+CL})-\hbar{2\pi\over 3}\psi (r_{+CL}) \label{eq: 149} \end{equation} From (\req{eq: 48}) we get: \begin{equation} \psi (r_{+CL}) = -\ln g_{CL}(r_{+CL})-{2\over \beta_{CL}} \int^L_{r_{+}}{dr\over g_{CL}(r)}-2\ln \left( {\beta_{CL}\over z_0}\right) \label{eq: 150} \end{equation} Using \req{eq: 131} for $g_{CL}$, \req{eq: 136} for $\beta_{CL}$, integrating the middle term and simplifying yields: \begin{eqnarray} \psi (r_{+CL}) &=& -{r_{-CL}\over r_{+CL}}\ln \left[{(r_{+CL}-r_{-CL})(L+r_{-CL})\over (r_{+CL}+r_{-CL}) (L-r_{-CL})}\right]+\ln({r_{+CL}^2-r^2_{-CL}\over r^2_{+CL}})\nonumber \\ &&\quad - \ln\left({L-r_{+CL}\over L+r_{+CL}}\right) +\ln\left({\Lambda z_0^2\over 8}\right) \label{eq: 151} \end{eqnarray} So the complete quantum corrected entropy is obtained by substituting \req{eq: 151} for $\psi (r_{+CL}$) and $m(r_{+CL})$ via \req{eq: 137} back into \req{eq: 149}. For large $L$ this result reduces to (subtracting off the constant term): \begin{eqnarray} S &=& S_{CL}+2\pi\beta_{CL}m(r_{+CL})+{\hbar 2\pi\over 3} {r_{-CL}\over r_{+CL}}\ln\left({r_{+CL}-r_{-CL}\over r_{+CL}+r_{-CL}}\right)\nonumber\\ &&\quad - {\hbar 2\pi\over 3}\ln\left({r_{+CL}^2-r^2_{-CL}\over r^2_{+CL}}\right) \label{eq: 152} \end{eqnarray} For an uncharged black hole then the entropy is given by: \begin{equation} S = S_{CL} + \hbar {2\pi\over 3}\ln\left[{L-r_{+CL}\over L+r_{+CL}}\right] \label{eq: 153} \end{equation} Note that using \req{eq: 141} the $m(r_{+CL})$ term reduces to a constant which we subtract off. Finally, in the extremal black hole limit \begin{equation} S = S_{CL}+2\pi\beta_{CL}m(r_{+CL}) \label{eq: 154} \end{equation} where $m(r_{+CL}) = {2\over 9}\hbar\Lambda r_{+CL}$ in the extremal case however $\beta_{CL}\rightarrow\infty $ so the entropy is divergent in this limit. \section{Conclusions} We have calculated the one-loop quantum corrections for generic dilaton gravity coupled to an Abelian gauge field. Both corrections to the black hole geometry and black hole thermodynamics were studied in detail. We then applied our generic results to the special cases of charged black holes in spherically symmetric gravity and rotating BTZ black holes. The former case enabled us to verify our results by comparison with the tree-level calculations of Braden et al. \cite{brown} and the one-loop corrections of Frolov et al. \cite{frolov}. Study of BTZ black holes is of particular interest due to recent revelations of a possible connection between string inspired black holes and BTZ geometry \cite{near}. Although our quantum corrected results can in principle be integrated exactly, numerical analysis will be required for rigourous study of particular theories. Such an analysis is in progress. Our hope is that ultimately such studies will lead to a better understanding of quantum thermodynamical processes associated with black holes and hence insight into the deep mysteries surrounding quantum gravity. \section{Acknowledgements} \par This work was supported in part by the Natural Sciences and Engineering Research Council of Canada. G.K. would like to thank J. Gegenberg for helpful conversations. We are also grateful to S.D. Odintsov for useful comments on the original version of the manuscript and for bringing several important references to our attention. \par\vspace*{20pt}
\section{Introduction} \label{s:int} Suppose we want to transmit or store a block of $l$ qubits (i.e. two-state quantum systems) in a noisy environment. Here `noisy' means that each qubit may become entangled with the environment. Thus due to spurious interactions with the environment the actual state of the $l$ qubits, described by a density operator $\rho(t)$, will differ from the original state $\ket{\Psi}$. This deviation can be quantified by the fidelity \begin{equation} F(t)=\bra{\Psi}\rho(t)\ket{\Psi} =1-\epsilon(t). \end{equation} In order to maximize this fidelity we may try all sorts of tricks ranging from the most obvious one i.e. isolating the qubits from the environment to more sophisticated methods such as ``symmetrisation" \cite{Deutsch1993,Barenco etal}, ``purification" \cite{Bennett etal,QPA}, and ``quantum error correction" \cite{Shor}. The last method seems to be the most popular one at the moment and relies on encoding the state of $l$ qubits into a set of $n$ qubits and trying to disentangle a certain number of qubits from the environment after some period of time. In the following we describe, very briefly, how some of these techniques work. We will assume that in the block of $l$ qubits each qubit is coupled to a different environment. This is a perfectly reasonable assumption, which is valid if the coherence length of the environment/reservoir is less than the spatial separation between the qubits~\cite{palma}, and introduces a great deal of simplifications to the calculations. Basically it allows us to view any dissipation of $l$ qubits as a set of independent dissipations of $l$ single qubits (i.e. we ignore collective phenomena such as superradiance etc.). The qubit--environment interaction leads to the qubit--environment entanglement, which in its most general form is given by \begin{eqnarray} \ket{0}\ket{R} & \longrightarrow & \ket{0}\ket{R_{00}(t)} + \ket{1}\ket{R_{01}(t)},\\ \ket{1}\ket{R} & \longrightarrow & \ket{0}\ket{R_{10}(t)} + \ket{1}\ket{R_{11}(t)}, \label{dissipation} \end{eqnarray} where states of the environment $\ket{R}$ and $\ket{R_{ij}}$ are neither normalised nor orthogonal to each other (thus we have to take additional care at the end of our calculations and normalise the final states). The r.h.s. of the formulae above can also be written in a matrix form as \begin{equation} \left( \begin{array}{cc} \ket{R_{00}} & \ket{R_{01}} \\ \ket{R_{10}} & \ket{R_{11}} \end{array} \right) \left( \begin{array}{c} \ket{0}\\ \ket{1}\end{array}\right ), \end{equation} and the 2 $\times$ 2 matrix can can be subsequently decomposed into some basis matrices e.g. into the unity and the Pauli matrices \begin{equation} \ket{R_0}1+\ket{R_1}\sigma_x + i\ket{R_2}\sigma_y + \ket{R_3}\sigma_z, \end{equation} where $\ket{R_0}= (\ket{R_{00}}+\ket{R_{11}})/2$, $\ket{R_3}= (\ket{R_{00}}-\ket{R_{11}})/2$, $\ket{R_1} = (\ket{R_{01}}+\ket{R_{10}})/2$, and $\ket{R_2} =(\ket{R_{01}}-\ket{R_{10}})/2$. Thus the qubit initially in state $\ket{\Psi}$ will evolve as \begin{equation} \ket{\Psi}\ket{R} \longrightarrow \sum_{i=0}^3\sigma_i\ket{\Psi}\ket{R_i} \label{action} \end{equation} becoming entangled with the environment (we have relabelled the unity operator and the Pauli matrices $\{1, \sigma_x, \sigma_y, \sigma_z\}$ respectively as $\{\sigma_0, \sigma_1, \sigma_2, \sigma_3\}$). Its fidelity with respect to the initial state $\ket{\Psi}$ evolves as \begin{equation} F(t) = \sum_{i,j} \bra{\Psi}\sigma_i\ket{\Psi} \bra{\Psi}\sigma_j\ket{\Psi}\bra{R_j(t)}R_i(t)\rangle . \label{fidelity} \end{equation} The formula (\ref{action}) describes how the environment affects any quantum state of a qubit and shows that a general qubit--environment interaction can be expressed as a superposition of unity and Pauli operators acting on the qubit. As we will see in the following, in the language of error correcting codes this means that the qubit state is evolved into a superposition of an error-free component and three erroneous components, with errors of the $\sigma_x$, $\sigma_y$ and $\sigma_z$ type. We can carry on this description even if the qubit itself is not in a pure state $\ket{\Psi}$ but is entangled with some other qubits. For example, if in a three qubit register initially in state $\ket{\tilde\Psi} = \ket{0}\ket{0}\ket{0}-\ket{1}\ket{1}\ket{1}$ the second qubit interacted with its environment then the state of the register at some time $t$ is given by \begin{equation} \sum_{i=0}^3\sigma_i^{(2)} \ket{\tilde\Psi}\ket{R_i (t)} = \sum_{i=0}^3(\ket{0}(\sigma_i\ket{0})\ket{0} - (\ket{1}(\sigma_i\ket{1})\ket{1})\ket{R_i (t)}, \label{int} \end{equation} where the superscript $(2)$ reminds us that the Pauli operators act only on the second qubit. We can then say that the second qubit was affected by quantum errors which are represented by the Pauli operators $\sigma_i$. Errors affecting classical bits can only change their binary values ($0\leftrightarrow 1$), in contrast quantum errors operators $\sigma_i$ acting on qubits can change their binary values ($\sigma_x$), their phases ($\sigma_z$) or both ($\sigma_y$). In general, a batch of $n$ qubits initially in some state $\ket{\tilde\Psi}$, each of them interacting with different environments, will evolve as \begin{equation} \prod_{k=1}^n\sum_{i=0}^3\sigma_i^{(k)} \ket{\tilde\Psi}\ket{R_i^{(k)} (t)}\;, \end{equation} namely multiple errors of the form $\sigma_i\otimes\sigma_j\cdot\cdot\cdot\otimes\sigma_k$ may occur, affecting several qubits at the same time. So much about unwelcome dissipation, what about remedies? \section{Stabilization via symmetrisation} The first proposed remedy was based on a symmetrisation procedure \cite{Deutsch1993}. The basic idea is as follows. Suppose you have a quantum system, you prepare it in some initial state $\ket{\Psi_i}$ and you want to implement a prescribed unitary evolution $\ket{\Psi (t)}$ or simply you want to preserve $\ket{\Psi_i}$ for some period of time $t$. Now, suppose that instead of a single system you can prepare $R$ copies of $\ket{\Psi_i}$ and subsequently you can project the state of the combined system on the symmetric subspace i.e. the subspace containing all states which are invariant under any permutation of the sub-systems. The claim is that frequent projections on the symmetric subspace will reduce errors induced by the environment. The intuition behind this concept is based on the observation that a prescribed error-free storage or evolution of the $R$ independent copies starts in the symmetric sub-space and should remain in that sub-space. Therefore, since the error-free component of any state always lies in the symmetric subspace, upon successful projection it will be unchanged and part of the error will have been removed. Note however that the projected state is generally not error--free since the symmetric subspace contains states which are not of the simple product form $\ket{\psi}\ket{\psi}\ldots\ket{\psi}$. Nevertheless it has been shown that the error probability will be suppressed by a factor of $1/R$ \cite{Barenco etal}. We illustrate here this effect in the simplest case of two qubits. The projection into the symmetric subspace is performed in this case by introducing the symmetrisation operator: \begin{equation} \label{symop2} S= \frac{1}{2} (P_{12}+P_{21})\;, \end{equation} where $P_{12}$ represents the identity and $P_{21}$ the permutation operator which exchanges the states of the two qubits. The symmetric--projection of a pure state $\ket{\Psi} $ of two qubits is just $S\ket{\Psi}$, which is then renormalised to unity. It follows that the induced map on mixed states of two qubits (including renormalisation) is: \begin{equation} \label{symproj2} \rho_1 \otimes \rho_2 \longrightarrow \frac{S(\rho_1 \otimes\rho_2 ) S^{\dagger}}{\mbox{Tr}{S(\rho_1 \otimes\rho_2 ) S^{\dagger}}} \end{equation} The state of either qubit separately is then obtained by partial trace over the other qubit. Consider for example the symmetric projection of $\rho\otimes \rho$ followed by renormalisation and partial trace (over either qubit) to obtain the final state ${\rho_s}$ of one qubit, given that the symmetric-projection was successful. A direct calculation based on (\ref{symproj2}) yields: \begin{equation} \label{twid} \rho \mapsto {\rho_s}= \frac{\rho +\rho^2}{\mbox{Tr}{(\rho +\rho^2 ) }} \end{equation} For any mixed state $\xi$ of a qubit the expression $\mbox{Tr}{\xi ^2 }$ provides a measure of the purity of the state, ranging from $\frac{1}{4}$ for the completely mixed state $I/2$ (where $I$ is the unit operator) to 1 for any pure state. From (\ref{twid}) we get \begin{equation} \mbox{Tr}{{\rho_s}^2} > \mbox{Tr}{\rho ^2} \end{equation} so that ${\rho_s}$ is {\em purer} than $\rho$. This illustrates that successful projection of a mixed state into the symmetric subspace tends to enhance the purity of the individual systems. To be more specific, let us assume now that the two copies are initially prepared in pure state $\rho_0 = \proj{\Psi}$ and that they interact with independent environments. After some short period of time $\delta t$ the state of the two copies $\rho^{(2)}$ will have undergone an evolution \begin{equation} \label{dec} \rho^{(2)}(0) = \rho_0 \otimes \rho_0 \hspace{5mm} \longrightarrow \hspace{5mm} \rho^{(2)}(\delta t) = \rho_1\otimes\rho_2 \end{equation} where $\rho_i = \rho_0 + \varrho_i$ for some Hermitian traceless $\varrho_i$. We will retain only terms of first order in the perturbations $\varrho_i$ so that the overall state at time $\delta t$ is \begin{eqnarray} \rho^{(2)} = \rho_0 \otimes\rho_0 + \varrho_1 \otimes \rho_0 +\rho_0 \otimes \varrho_2 + O(\varrho_1 \varrho_2 )\;. \label{decst} \end{eqnarray} We can calculate the average purity of the two copies before symmetrisation by calculating the average trace of the squared states: \begin{equation} \label{this} \frac{1}{2} \sum_{i=1}^2 \mbox{Tr}( (\rho_0+\varrho_i)^2) = 1 + 2 \mbox{Tr}(\rho_0\tilde\varrho ), \end{equation} where $\tilde\varrho=\frac{1}{2}(\varrho_1+\varrho_2)$. Note that $\mbox{Tr}({\rho_0 \tilde\varrho})$ is negative, so that the expression above does not exceed 1. After symmetrisation each qubit is in state \begin{equation} \rho_s=[1- \mbox{Tr}(\rho_0\tilde\varrho )] \rho_0 +\frac{1}{2}\tilde\varrho +\frac{1}{2}(\rho_0\tilde\varrho+ \tilde\varrho\rho_0) \end{equation} and has purity \begin{equation} \mbox{Tr}(\rho_s^2)=1 + \mbox{Tr}(\rho_0\tilde\varrho). \end{equation} Since $\mbox{Tr}{\rho_s^2 }$ is closer to 1 than (\ref{this}), the resulting symmetrised system $\rho_s$ is left in a purer state. Let us now see how the fidelity changes by applying the symmetrisation procedure. The average fidelity before symmetrisation is \begin{equation} \label{fid1} F_{bs}=\frac{1}{2} \sum_i \bra{\Psi}\rho_0 +\varrho_i\ket{\Psi} = 1+ \bra{\Psi} \tilde\varrho \ket{\Psi}\;, \end{equation} while after successful symmetrisation it takes the form \begin{equation} F_{as}=\bra{\Psi}\rho_s \ket{\Psi} = 1+ \frac{1}{2} \bra{\Psi}\tilde\varrho \ket{\Psi}\;. \end{equation} The state after symmetrisation is therefore closer to the initial state $\rho_0$. For the generic case of $R$ copies the purity of each qubit after symmetrisation is given by \cite{Barenco etal} \begin{equation} \mbox{Tr}(\rho_s^2)=1 + 2\frac{1}{R} \mbox{Tr}(\rho_0\tilde\varrho)\;, \label{purR} \end{equation} where now $\tilde\varrho=\frac{1}{R}\sum_{i=1}^R\varrho_i$, and the fidelity takes the form \begin{equation} \label{fid2} \bra{\Psi}\rho_s \ket{\Psi} = 1+ \frac{1}{R} \mbox{Tr} ({\rho_0 \tilde\varrho })\;. \end{equation} Formulae (\ref{purR}) and (\ref{fid2}) must be compared with the corresponding ones before symmetrisation, i.e. (\ref{this}) and (\ref{fid1}). As we can see, $\rho_s$ approaches the unperturbed state $\rho_0$ as $R$ tends to infinity. Thus by choosing $R$ sufficiently large and the rate of symmetric projection sufficiently high, the residual error at the end of a computation can, in principle, be controlled to lie within any desired small tolerance. The efficiency of the symmetrisation procedure depends critically on the probability that the state of the $R$ qubits is successfully projected into the symmetric subspace. It has been shown that if the projections are done frequently enough, then the cumulative probability that they all succeed can be made as close as desired to unity. This is a consequence of the fact that the fidelity of the state of the $R$ computers with respect to the corresponding error free state for small times $\delta t$ has a parabolic behaviour (see section \ref{s:dyn}). Therefore the probability of successful projection, which is unity at the initial time, begins to change only to second order in time. If we project $n$ times per unit time interval, i.e. we choose the time interval between two subsequent projections $\delta t = 1/n$, then the cumulative probability that all projections in one unit time interval succeed is given by \begin{eqnarray} [1-k(\delta t)^2 ]^n = (1-\frac{k}{n^2})^n \rightarrow 1 \mbox{ as } n\rightarrow \infty\;. \end{eqnarray} Here $k$ is a constant depending on the rate of rotation of the state out of the symmetric subspace. This effect is known as the ``quantum watch-dog effect'' or the ``quantum Zeno effect". \section{Quantum encoding and decoding} \label{s:enc-dec} The idea of protecting information via encoding and decoding lies at the foundations of the classical information theory. It is based on a clever use of redundancy during the data storage or transmission. For example, if the probability of error (bit flip) during a single bit transmission via a noisy channel is $p$ and each time we want to send bit value 0 or 1 we can encode it by a triple repetition i.e. by sending 000 or 111. At the receiving end each triplet is decoded as either zero or one following the majority rule - more zeros means 0, more ones means 1. This is the simplest error correcting protocol which allows to correct up to one error. In the triple repetition code the signalled bit value is recovered correctly both when there was no error during the transmission of the three bits, which happens with probability $(1-p)^3$, and when there was one error at any of the three locations, which happens with probability $3p(1-p)^2$. Thus the probability of the correct transmission (up to the second order in $p$) is $1-3p^2$ i.e. the probability of error is now $3p^2$, which is much smaller when compared with the probablity of error without encoding and decoding $p$ ($p\ll 1$). This way we can trade the probability of error in the signalled message for a number of transmissions via the channel. In our example the reduction of the error rate from $p$ to $3p^2$ required to send three times more bits. If sending each bit via the channel costs us money we have to decide what we treasure more, our bank account or our infallibility. The triple repetition code encodes one bit into three bits and protects against one error, in general we can construct codes that encode $l$ bits into $n$ bits and protect against $t$ errors. The best codes, of course, are those which for a fixed value $l$ minimize $n$ and maximize $t$. Quantum error correction which protects quantum states is a little bit more sophisticated simply because the bit flip is not the only ''quantum error" which may occur, as we have seen in the previous sections. Moreover, the decoding via the majority rule does not usually work because it may involve measurements which destroy quantum superpositions. Still, the triple repetition code is a good starting point to investigate quantum codes and even to construct the simplest ones. The simplest interesting case of the most general qubit--environment evolution (\ref{dissipation}) is the case of decoherence \cite{Zurek1991} where the environment effectively acts as a measuring apparatus \begin{eqnarray} \ket{0}\ket{R} & \longrightarrow & \ket{0}\ket{R_{00}(t)},\\ \ket{1}\ket{R} & \longrightarrow & \ket{1}\ket{R_{11}(t)}. \label{decoherence} \end{eqnarray} Following our discussion in Section \ref{s:int} we can see that this model leads only to dephasing errors of the $\sigma_z$ type. It turns out that a phase flip can be handled almost in the same way as a classical bit flip. Again, consider the following scenario: we want to store, in a computer memory, one qubit in an {\em unknown} quantum state of the form $\alpha\ket{0}+\beta\ket{1}$ and we know that any single qubit which is stored in a register may, with a small probability $p$, undergo a decoherence type entanglement with an environment (Eq. \ref{decoherence}); in particular \begin{equation} (\alpha\ket{0}+\beta\ket{1})\ket{R}\longrightarrow \alpha\ket{0}\ket{R_{00}}+\beta\ket{1}\ket{R_{11}}. \end{equation} Let us now show how to reduce the probability of decoherence to be of the order $p^2$. Before we place the qubit in the memory register we {\em encode} it: we can add two qubits, initially both in state $\ket{0}$, to the original qubit and then perform an encoding unitary transformation \begin{eqnarray} \ket{000}&\longrightarrow &\ket{C_0} =(\ket{0}+\ket{1})(\ket{0}+\ket{1})(\ket{0}+\ket{1}),\\ \ket{100}&\longrightarrow &\ket{C_1} =(\ket{0}-\ket{1})(\ket{0}-\ket{1})(\ket{0}-\ket{1}), \end{eqnarray} generating state $\alpha\ket{C_0}+\beta\ket{C_1}$. Now, suppose that only the second stored qubit was affected by decoherence and became entangled with the environment: \begin{eqnarray} \alpha (\ket{0}+\ket{1})(\ket{0}\ket{R_{00}} +\ket{1}\ket{R_{11}})(\ket{0}+\ket{1}) +\nonumber\\ \beta (\ket{0}-\ket{1})(\ket{0}\ket{R_{00}} -\ket{1}\ket{R_{11}})(\ket{0}-\ket{1}), \end{eqnarray} which, following Eq. (\ref{int}), can be written as \begin{equation} (\alpha\ket{C_0} + \beta\ket{C_1})\ket{R_0} + \sigma_z^{(2)}(\alpha\ket{C_0} + \beta\ket{C_1})\ket{R_3}. \end{equation} If vectors $\ket{C_0}$, $\ket{C_1}$, $\sigma_z^{(k)}\ket{C_0}$, and $\sigma_z^{(k)}\ket{C_1}$ are orthogonal to each other we can try to perform a measurement on the qubits and project their state either on the state $\alpha\ket{C_0} + \beta\ket{C_1}$ or on the orthogonal one $\sigma_z^{(2)}(\alpha\ket{C_0} + \beta\ket{C_1})$. The first case yields the proper state right away, the second one requires one application of $\sigma_z$ to compensate for the error. In this simple case one can even find a direct unitary operation which can fix all one qubit phase flips regardless their location. For example the transformation \begin{eqnarray} \ket{000}\to \ket{000} & & \ket{100} \to \ket{011}\nonumber \\ \ket{001}\to \ket{001} & & \ket{101} \to \ket{110}\nonumber \\ \ket{010}\to \ket{010} & & \ket{110} \to \ket{101}\nonumber \\ \ket{011}\to \ket{111} & & \ket{111} \to \ket{100} \end{eqnarray} corrects any single bit flip $0\leftrightarrow 1$ and when applied in the conjugate basis ($\ket{0'}=\ket{0}+\ket{1}$, $\ket{1'}=\ket{0}-\ket{1}$) it corrects any single phase flip (the bit flips become phase flips in the new basis). The snag is that using the scheme above we can correct up to one phase error $\sigma_z$ or we can go to a conjugate basis and the same scheme will correct up to one amplitude error $\sigma_x$ but it cannot correct up to one general error, be it amplitude or phase. To fix this problem Peter Shor in 1995 combined the phase and the amplitude correction schemes into one constructing the following nine qubit code \cite{Shor}: \begin{eqnarray} &&\ket{0}\to \frac{1}{2\sqrt{2}}(\ket{000}+\ket{111})(\ket{000}+\ket{111}) (\ket{000}+\ket{111}) \label{shor1}\\ &&\ket{1}\to \frac{1}{2\sqrt{2}}(\ket{000}-\ket{111})(\ket{000}-\ket{111}) (\ket{000}-\ket{111})\;. \label{shor2} \end{eqnarray} This code involves double encoding, first in base $\ket{0}$ and $\ket{1}$ and then in base $\ket{0'}$ and $\ket{1'}$, and it allows to correct up to one either bit or phase flip. It turns out that the ability to correct both amplitude and phase errors suffices to correct any error due to entanglement with the environment. In other words the action of the environment on qubits can be viewed in terms of bit and phase flips. \section{Quantum error-correcting codes} The original nine qubit code of Shor can be further simplified. It has been shown that a five qubit code suffices to correct a single error of any type. Let us now specify the conditions for the existence of quantum error-correcting codes. We say we can correct a single error $\sigma_i^{(k)}$ (where $i=0\ldots 3$ refers to the type of error) if we can find a transformation such that it maps all states with a single error $\sigma_i^{(k)}\ket{\tilde\Psi}$ into the proper error free state $\ket{\tilde\Psi}$: \begin{equation} \sigma^{(k)}_i \ket{\tilde\Psi} \longrightarrow \ket{\tilde\Psi} \end{equation} To make it unitary we may need an ancilla \begin{equation} \sigma_i^{(k)} \ket{\tilde\Psi}\ket{0} \longrightarrow \ket{\tilde\Psi} \ket{a_i^k}\;. \end{equation} For encoded basis states of a single qubit $\ket{C_0}$ and $\ket{C_1}$ this implies \cite{bdws} \begin{eqnarray} A_k \ket{C_0 }\ket{0} & \longrightarrow & \ket{C_0}\ket{a_k}\\ A_k \ket{C_1 }\ket{0} & \longrightarrow & \ket{C_1}\ket{a_k}, \end{eqnarray} where $A_k$ denotes all the possible types of independent errors affecting at most one of the qubits.The above requirement leads to the following unitarity conditions \begin{eqnarray} \bra{C_0} A^\dagger_k A_l\ket{C_0}&=&\bra{C_1} A^\dagger_k A_l\ket{C_1} = \bra{a_k} a_l\rangle\;, \label{condec}\\ \bra{C_0}A^\dagger_k A_l\ket{C_1}&=& 0 \;. \end{eqnarray} The above conditions are straightforwardly generalised to an arbitrary $t$ error correcting code, which corrects any kind of transformations affecting up to $t$ qubits in the encoded state. In this case the operators $A_k$ are all the possible independent errors affecting up to $t$ qubits, namely operators of the form $\Pi_{i=1}^t\sigma_i$ acting on $t$ different qubits. In the case of the so-called ``nondegenerate codes'' Eq. (\ref{condec}) takes the simple form \cite{nostro} \begin{equation} \bra{C_0} A^\dagger_k A_l\ket{C_0} =\bra{C_1} A^\dagger_k A_l\ket{C_1}=0\;. \label{orth} \end{equation} This condition requires that all states which are obtained by affecting up to $t$ qubits in the encoded states are all orthogonal to each other, and therefore distinguishable. This ensures that by performing suitable projections of the encoded state we are able to detect the kind of error which occurred and ``undo'' it to recover the desired error free state. Condition (\ref{orth}), even if more restrictive than (\ref{condec}), is particularly useful because it allows to establish bounds on the resources needed in order to have efficient nondegenerate codes. Let us assume that the initial state of $l$ qubits is encoded in a redundant Hilbert space of $n$ qubits. If we want to encode $2^l$ input basis states and correct up to $t$ errors we must choose the dimension of the encoding Hilbert space $2^n$ such that all the necessary orthogonal states can be accomodated. According to Eq. (\ref{orth}), the total number of orthogonal states that we need in order to be able to correct $i$ errors of the three types $\sigma_x$, $\sigma_y$ and $\sigma_z$ in an $n$-qubit state is $3^i\left(\begin{array}{c} n \\ i \end{array}\right)$ (this is the number of different ways in which the errors can occur). The argument based on counting orthogonal states then leads to the following condition \begin{eqnarray} 2^l\sum_{i=0}^t 3^i\left(\begin{array}{c} n \\ i \end{array}\right)\leq 2^n.\label{hamming} \end{eqnarray} Eq. (\ref{hamming}) is the quantum version of the Hamming bound for classical error-correcting codes~\cite{macw}; given $l$ and $t$ it provides a lower bound on the dimension of the encoding Hilbert space for nondegenerate codes. Let us mention that an explicit construction for quantum codes for some values $(l,n,t)$ which saturate the quantum Hamming bound has been provided \cite{gottesman}. It is interesting that this bound has not been beaten so far by degenerate codes \cite{barolo}. The quantum version of the classical Gilbert-Varshamov bound~\cite{macw} can be also obtained, which gives an upper bound on the dimension of the encoding Hilbert space for optimal non degenerate codes: \begin{eqnarray} 2^l\sum_{i=0}^{2t} 3^i\left(\begin{array}{c} n \\ i \end{array}\right)\geq 2^n.\label{gvbound} \end{eqnarray} This expression can be proved from the observation that in the $2^n$ dimensional Hilbert space with a maximum number of encoded basis vectors (or code-vectors) ${\left | \, C^k \right \rangle}$ any vector which is orthogonal to ${\left | \, C^k \right \rangle}$ (for any $k$) can be reached by applying up to $2t$ error operations of $\sigma_x$, $\sigma_y$, and $\sigma_z$ type to any of the $2^l$ encoded basis vectors. Clearly all vectors which cannot be reached in the $2t$ operations can be added to the encoded basis states ${\left | \, C^k \right \rangle}$ as all the vectors into which they can be transformed by applying up to $t$ amplitude and/or phase transformations are orthogonal to all the others. This situation cannot happen because we have assumed that the number of code-vectors is maximal. Thus the number of orthogonal vectors that can be obtained by performing up to $2t$ transformations on the code-vectors must be at least equal to the dimension of the encoding Hilbert space. It follows from Eq. (\ref{hamming}) that a one-bit quantum error correcting code to protect a single qubit ($l=1$, $t=1$) requires at least $5$ encoding qubits and, according to Eq. (\ref{gvbound}), this can be achieved with less than $10$ qubits. Indeed, Shor's nine qubit code can be simplified to the seven qubit code \cite{steane}, and ultimately to the quantum Hamming bound \cite{bdws,lafl}. We will consider explicitly one form of the five qubit code in Section \ref{s:beyond}. The asymptotic form of the quantum Hamming bound (\ref{hamming}) in the limit of large $n$ is given by \begin{eqnarray} \frac{l}{n}\leq 1-\frac{t}{n}\log_2 3 -H(\frac{t}{n}),\label{hamasym} \end{eqnarray} where $H$ is the entropy function $H(x)=-x \log_2 x-(1-x)\log_2(1-x)$. The corresponding asymptotic form for the quantum Gilbert-Varshamov bound (\ref{gvbound}) is \begin{equation} \frac{l}{n}\geq 1-\frac{2t}{n}\log_2 3 -H(\frac{2t}{n}).\label{gvasym} \end{equation} As we can see from eq. (\ref{hamasym}), in quantum error correction there is an upper bound on the error rate $t/n$ which a code can tolerate. In fact, differently from the classical case, where any arbitrary error rate can be corrected by a suitable code, in the quantum world the ratio $t/n$ cannot be larger than 0.18929 for nondegenerate codes. \section{System-environment dynamics} \label{s:dyn} In order to provide a tangible illustration of some abstract ideas discussed in the text we have picked up the most popular quantum-optical model of dissipation commonly used to describe spontaneous emission. A two level atom, with two energy eigenstates $\ket{0}$ and $\ket{1}$ separated by $\hbar\omega_0$, interacting with an environment modelled as a set of quantised harmonic oscillators, e.g. a set of quantised modes of radiation with frequencies $\omega_m$. The Hamiltonian of the combined system $H= H_0 +V$ includes both the free evolution of the qubit and the environment. The free evolution Hamiltonian is given by \begin{equation} H_0 = \hbar\omega_0 \proj{1} +\sum_m \hbar\omega_m a^\dagger_m a_m\;, \end{equation} where $a_m$ and $a^\dagger_m$ represent the annihilation and creation operators of the radiation mode of frequency $\omega_m$. The interaction (in the rotating wave approximation) is described by \begin{equation} V = \sum_m \lambda_m \ket{0}\bra{1} a^\dagger_m + \lambda^\star_m \ket{1}\bra{0} a_m\;, \end{equation} where $\lambda_m$ is the coupling constant between the qubit and the mode of frequency $\omega_m$. In order to find the time evolution of the relative states of the environment $\ket{R_i(t)}$ we need some knowledge about the qubit-environment interaction. Let us then have a closer look at a dissipative dynamics in our model of a qubit coupled to a continuum of field modes or harmonic oscillators. If all the oscillators in the environment are in their ground states and the qubit is initially prepared in state $\ket{\Psi}=\alpha\ket{0}+\beta\ket{1}$ then the dynamics described by the Hamiltonian $H=H_0+V$ does not affect state $\ket{0}$. It is state $\ket{1}$ which undergoes a decay. Let us then consider a case when the initial state of the combined system (qubit+environment) is \begin{equation} \ket{\phi_i}=\ket{1}(\ket{0}_1\ket{0}_2\ldots\ket{0}_f\ldots\ket{0}_{max}), \end{equation} meaning the qubit is in state $\ket{1}$ and all the harmonic oscillators in their ground states $\ket{0}$ (we will denote the state where all harmonic oscillators are in the ground state as the vacuum $\ket{\bf 0}$). Possible final states of the combined system are \begin{equation} \ket{\phi_f}=\ket{0}(\ket{0}_1\ket{0}_2\ldots\ket{1}_f\ldots\ket{0}_{max}), \end{equation} where the qubit decayed to state $\ket{0}$ and one of the harmonic oscillators got excited. Let us note that \begin{equation} H_0 \ket{\phi_i} = \hbar\omega_0 \ket{\phi_i} ,\quad H_0 \ket{\phi_f} = \hbar\omega_f \ket{\phi_f},\quad \bra{\phi_f} H_0 \ket{\phi_i} = 0 , \quad \bra{\phi_f} V \ket{\phi_i} = \lambda_f . \label{Ham} \end{equation} Let us write $\ket{\phi (t)}$ as \begin{equation} \ket{\phi (t)} = c_i(t)e^{-i\omega_0 t}\ket{\phi_i} + \sum_f c_f (t) e^{-i\omega_f t}\ket{\phi_f} \end{equation} which, using our notation from the previous section, implies $\ket{R_{00}}=\ket{\bf{0}}$, $\ket{R_{01}}=0$, $\ket{R_{10}(t)}=\sum_f c_f(t)e^{-i\omega_ft}\ket{1_f}$ and $\ket{R_{11}(t)}=c_i(t)e^{-i\omega_0 t}\ket{\bf{0}}$. In order to find the relevant time dependance we have to solve the Schr\"odinger equation \begin{eqnarray} i\hbar \dot{c}_i (t) & =& \sum_f \lambda_f^\star e^{-i(\omega_f - \omega_0) t} c_f (t) \label{Schr1}\\ i\hbar \dot{c}_f (t) & =& \lambda_f e^{i(\omega_f - \omega_0) t} c_i (t). \label{Schr2} \end{eqnarray} The second equation can be solved formally for $c_f (t)$ \begin{equation} c_f(t) = - \frac{i}{\hbar} \int_0^t dt' \lambda_f e^{i(\omega_f - \omega_0) t'} c_i (t') \label{eqcf} \end{equation} and after substituting this expression for $c_f(t)$ in Eq.(\ref{Schr1}) we obtain \begin{equation} \dot {c}_i(t) = - \int_0^t dt' K(t-t') c_i(t'), \qquad K(\tau) = \frac{1}{\hbar^2}\sum_f|\lambda_f|^2 e^{-i(\omega_f - \omega_0) \tau}. \end{equation} It is the function $\lambda_f = \lambda (\omega_f)$ which determines the character of the evolution. \begin{itemize} \item Parabolic Decay. At short times, the exponential in $e^{-i(\omega_f - \omega_0) (t-t')}$ in $K(t-t')$ can be replaced by $1$. This is justified when $t\ll \frac{1}{\Delta}$, where $\Delta$ is a typical width of the $\lambda (\omega_f)$ curve. Usually, for a bell-shaped $\lambda (\omega_f)$ curve the order of $\Delta$ is pretty well approximated by $\omega_0$. For example if we analyse spontaneous emission in the optical domain then $\omega_0 = \Delta = 10^{15}\mbox{Hz}$ thus the short time means here much less than $10^{-15}$ s. The integration in Eq. (\ref{eqcf}) together with the initial condition $c_i(t=0) = 1$ gives \begin{equation} |c_i(t)|^2 = |\bra{\phi_i}\phi(t)\rangle|^2 = 1 - 2\frac{t^2}{\hbar^2} \sum_f \lambda^2_f. \end{equation} The same result can be otained obtained directly by writing \begin{equation} \ket{\phi(t)} = e^{-iHt/\hbar}\ket{\phi_i} = (1 - \frac{i}{\hbar} Ht - \frac{1}{\hbar^2} H^2 t^2 + \ldots)\ket{\phi_i} \end{equation} which, together with Eq.(\ref{Ham}) gives \begin{equation} |\bra{\phi_i}\phi(t)\rangle|^2 = 1- 2\frac{t^2}{\hbar^2} (\langle H^2 \rangle - \langle H \rangle^2) \ldots = 1-2\frac{t^2}{\hbar^2}\sum_f \lambda^2_f + \ldots \end{equation} Thus for short times the decay is always parabolic. Let us mention in passing that from a purely mathematical point of view we have assumed here that expression $(\langle H^2\rangle -\langle H\rangle^2) = \sum_f \lambda^2_f $, i.e. the variance of the energy in the initial state $\ket{\phi_i}$, is finite. Needless to say in reality it is always finite but there are mathematical models in which, due to various approximations, this may not be the case (e.g. the Lorentzian distribution which has no finite moments). \item Exponential Decay. Expression $|\lambda_f|^2 e^{-i(\omega - \omega_0) \tau}$ viewed as a function of $\omega_f -\omega_0$ oscillates with frequency $1/\tau$ whereas $\lambda_f = \lambda (\omega_f)$ varies smoothly in the frequency domain. Again taking $\Delta$ as the typical width of the $\lambda (\omega_f)$ curve for $\tau >> 1/ \Delta$ the sum in $K(\tau)$ averages out to zero. This allows to substitute $c_i(t)$ for $c_i(t')$ in Eq.(\ref{Schr1}) which gives \begin{equation} \dot {c}_i(t) \approx -c_i(t)\int_0^t d\tau K(\tau) \approx -c_i(t)\int_0^\infty d\tau K(\tau). \end{equation} Now we can calculate $\int_0^\infty d\tau K(\tau)$ using the identity \begin{equation} \int_0^\infty d \tau e^{i\omega\tau} = \lim_{\epsilon \rightarrow 0^+} \int_0^\infty d \tau e^{i(\omega + i\epsilon)\tau} = \lim_{\epsilon \rightarrow 0^+}\frac{i}{\omega+i\epsilon} = i\mbox{\cal P}\frac{1}{\omega} + \pi \delta (\omega). \end{equation} It gives \begin{equation} \int_0^\infty d\tau K(\tau) = \frac{\gamma}{2} + i \delta, \quad \frac{\gamma}{2} = \frac{\pi}{\hbar^2} |\lambda (\omega_f=\omega_0)|^2, \quad \delta =\mbox{\cal P}\sum_f\frac{|\lambda_f|^2}{\omega_0-\omega_f}. \end{equation} Incorporating the energy shift $\hbar \delta$ into the modified energy separation $\hbar (\omega_0+\delta)$ we finally obtain \begin{equation} \dot{c}_i(t) = -\frac{\gamma}{2} c_i(t)\qquad \mbox{that is} \qquad c_i(t) = e^{-\frac{\gamma t}{2}} \end{equation} and consequently \begin{equation} c_f(t)=\frac{\lambda_f}{\hbar} \frac{1-e^{i(\omega_f-\omega'_0 + i\gamma/2)t}}{\omega_f-\omega'_0 + i\gamma/2} \end{equation} \end{itemize} Let us now go back to the language introduced in section \ref{s:int}. The states of the environment $\ket{R_0(t)},\\ \ket{R_1(t)}, \ket{R_2(t)}$ and $\ket{R_3(t)}$ in the present context take the explicit form \begin{eqnarray} \ket{R_0(t)}&=&\frac{1}{2}[1+c_i(t)e^{-i\omega_0 t}]\ket{\bf 0}\;, \label{R0}\\ \ket{R_1(t)}&=&\frac{1}{2}\sum_f c_f(t) e^{-i\omega_f t}\ket{1}_f\;, \label{R1}\\ \ket{R_2(t)}&=&-\frac{1}{2}\sum_f c_f(t) e^{-i\omega_f t}\ket{1}_f\;, \label{R2}\\ \ket{R_3(t)}&=&\frac{1}{2}[1-c_i(t)e^{-i\omega_0 t}]\ket{\bf 0}\;. \label{R3} \end{eqnarray} By formula (\ref{fidelity}), the fidelity of this process is given by \begin{eqnarray} F(t)&=&\bra{R_0(t)} R_0(t)\rangle + \bra{R_3(t)} R_3(t) \rangle -2\mbox{Re}\bra{R_0(t)}R_3(t)\rangle\nonumber\\ &=&|c_i(t)|^2\;. \end{eqnarray} Therefore, the fidelity in the case of a parabolic decay takes the form \begin{eqnarray} F_{par}(t)= 1 - 2\frac{t^2}{\hbar^2} \sum_f \lambda^2_f\;, \end{eqnarray} while in the case of an exponential decay it has the exponential form \begin{eqnarray} F_{exp}(t)= e^{-\gamma t}\;. \end{eqnarray} \section{Benefits of quantum error correction} \label{s:beyond} In order to get an idea about the efficiency of quantum error correction, we will now discuss an explicit construction of the single error-correcting five qubit code. The initial state of the qubit $\alpha\ket{0}+\beta\ket{1}$ is encoded in state $\alpha\ket{C_0}+\beta\ket{C_1}$, where \cite{lafl} \begin{eqnarray} \ket{C_0}&=&\ket{00010}+\ket{00101}-\ket{01011}+\ket{01100} +\ket{10001}-\ket{10110}-\ket{11000}-\ket{11111}\\ \ket{C_1}&=&\ket{00000}-\ket{00111}+\ket{01001}+\ket{01110} +\ket{10011}-\ket{10100}+\ket{11010}-\ket{11101}. \end{eqnarray} (To see the benefits of quantum error correction we do not need to use the explicit form of the code, we wrote it down here for those curious readers who may want to play with quantum error correcting codes.) These encoded states are chosen in such a way that conditions (\ref{orth}) are satisfied. Since this code can correct any type of error affecting one qubit, it is suitable for protecting quantum states against spontaneous emission. We notice that the spontaneous emission process described in Sect. \ref{s:dyn}, unlike decoherence, involves both phase and amplitude errors and therefore it cannot be successfully defeated with less than five bit codes. The probability that the state undergoes exponential decay in the presence of spontaneous emission is approximately given by \begin{eqnarray} P_{dec}(t)=1-F_{exp}(t)=1-e^{-\gamma t}\;. \end{eqnarray} If we assume that the five qubits decay independently from each other, the probability that none of them decays is given by \begin{eqnarray} P_{no\;dec}(t)=e^{-5\gamma t}\; \end{eqnarray} while the probability that only one of them decays is \begin{eqnarray} P_{1\;dec}(t)=e^{-4\gamma t}(1-e^{-\gamma t})\;. \end{eqnarray} Since by construction the above error correction scheme corrects perfectly the encoded state when only one of the qubits is affected, the fidelity of reconstruction of the state after the error correction is at least as high as the probability of having at most one qubit decay during the process, that is \begin{eqnarray} F_{ec}(t)\ge P_{no\;dec}(t)+ 5P_{1\;dec}(t)=e^{-4\gamma t}(5-4e^{-\gamma t}). \end{eqnarray} In order to have a successful error correction the such fidelity must be greater than the fidelity $F_{exp}(t)$ corresponding to a single qubit in the absence of error correction. This is true when the decay probability $P_{dec}(t)$ is much smaller than one, namely when the correction procedure is applied at times $t\ll 1/\gamma$. Actually, for $t\ll 1/\gamma$ the fidelity of reconstruction after error correction is bounded by \begin{eqnarray} F_{ec}(t)\ge 1-10\gamma^2 t^2 + O(t^3)\;, \end{eqnarray} namely it has parabolic form, while the single qubit decay probability is \begin{eqnarray} P_{dec}(t)\simeq 1-\gamma t\;. \end{eqnarray} \section{Concluding remarks} Research in quantum error correction in its all possible variations has become vigorously active and any comprehensive review of the field must be obsolete as soon as it is written. Here we have decided to provide only some very basic knowledge, hoping that this will serve as a good starting point to enter the field. The reader should be warned that we have barely scratched the surface of the current activities in quantum error correction neglecting topics such as group theoretical ways of constructing good quantum codes \cite{GF4}, concatenated codes \cite{knill}, quantum fault tolerant computation \cite{divi-shor} and many others. Many interesting papers in these and many related areas can be found at the Los Alamos National Laboratory e-print archive (http://xxx.lanl.gov/archive/quant-ph). This work was supported in part by the European TMR Research Network ERP-4061PL95-1412, the TMR Marie Curie Fellowship Programme, Hewlett-Packard, The Royal Society London and Elsag-Bailey, a Finmeccanica Company.
\section{Introduction} Stochastic partial differential equations (SPDEs) are an essential tool in modeling systems where noise is relevant~\cite{MSR}. SPDEs are used for models of many macroscopic systems, from turbulence~\cite{FNS,Frisch,Lvov-Procaccia}, to pattern-formation~\cite{KPZ,MHKZ}, to the structural development of the Universe itself~\cite{Berera-Fang,Hochberg-Mercader,PGHL,BDGP,Dominguez-et-al}. It is known that certain SPDEs can be studied with tools that transform them into equivalent (stochastic) field theories which exhibit deep and important relationships with quantum field theory (QFT). See, for example, \cite{MSR,FNS,KPZ,MHKZ} and~\cite{De-Dominicis-Peliti,Sun-Plischke,Frey-Tauber,Zinn-Justin}. In this paper we set up the field theoretical ``minimalist formalism'' for SPDEs, and demonstrate how to extract the one-loop physics for an {\em arbitrary} SPDE subject to additive Gaussian noise. It is important to realize that Gaussian noise does {\em not\,} imply that the field degrees of freedom undergo Gaussian fluctuations: the combined interplay of interactions and fluctuations will appear in the third (and higher) cumulants for the field $\phi(\vec x,t)$. Also, the limitation to one-loop physics is not as serious as might be supposed: Experience with quantum field theories (QFTs) has taught us that one-loop physics is often quite adequate to give a good description of the salient issues \cite{G-M+L,HMPV-tutorial,HMPV-a2,Weinberg-1,Weinberg-2,Rivers}. In fact, in QFT, the calculation of one-loop quantities can be augmented by means of ``Renormalization Group improved perturbation theory'', which contains most of the relevant features of the physics to {\em all\,} orders in the expansion parameter~\cite{HMPV-tutorial,HMPV-a2}. (This was called ``magical perturbation theory'' by the authors~\cite{G-M} of reference \cite{G-M+L}.) Furthermore, at one-loop (and higher), one can also introduce the effective action and effective potential~\cite{Perturbative,Non-perturbative,Coleman-Weinberg,Jackiw}, tools that allow one to determine the combined effects of interactions and fluctuations on the ground state of the system. Defining and calculating the one-loop effective action and effective potential is straightforward. Interpreting the physical significance of these quantities is more subtle. For arbitrary SPDEs it may not even be meaningful to define a notion of physical energy. Even when the physical energy makes sense, dissipative effects may vitiate energy conservation (even when noise is absent). We therefore spend some effort in establishing that certain key features of the effective action for QFTs carry over to SPDEs. In particular, we demonstrate that it is still meaningful to define and calculate the effective potential and look for its minima. The minima of the effective potential correspond to ground states of the system, and the locations of these minima are equal to stochastic expectation values of the fluctuating field in the presence of noise. While it is possible to provide an abstract non-perturbative definition of the effective action~\cite{Non-perturbative}, in order to proceed with explicit calculations (such as for the one-loop effective action) one needs a perturbative procedure based on an expansion in some small parameter. A well-known procedure of this type, the Martin--Siggia--Rose (MSR) formalism, already exists in the literature~\cite{MSR}. The MSR formalism invokes additional (unphysical) ``conjugate fields'', which are generalizations of the fictitious fields sometimes introduced to deal with the dynamics of diffusion. These fictitious fields permit one to extend some of the procedures of conservative physical systems to diffusion. For instance, Morse and Feshbach state: ``the dodge is to consider, $\dots$, a `mirror-image' system with negative friction, into which the energy goes which is drained from the dissipative system''. (See~\cite{Morse-Feshbach} page 298.) In this paper we do not make use of the conjugate field formalism of MSR and, instead, proceed in a direct way in which we only have physical fields, (plus possibly a nontrivial functional Jacobian that can be rewritten in terms of ghost fields). This approach simplifies the calculation since it halves the number of fields one has to deal with. These two {\em alternative} formalisms are very similar to the situation in spontaneously broken gauge field theories, where one can use two {\em equivalent\,} approaches to perturbation theory, such as ``unitary gauges'' versus ``renormalizable gauges'': in one case the {\em particle content\,} is explicit and in the other {\em renormalizability} is explicit. After setting up the path integral formalism for the characteristic functional (partition function), $Z[J]$, we define both the perturbative~\cite{Perturbative} and non-perturbative effective action~\cite{Non-perturbative}. We then focus on the one-loop effective action and its restriction to constant (homogeneous and stationary) fields: the effective potential~\cite{Weinberg-2,Rivers}. An important result is that the {\em amplitude} of the noise two-point correlation function acts as the loop-counting parameter and is the analog of Planck's constant $\hbar$ in this SPDE context. We conclude by deriving the formula for the one-loop effective potential of a general SPDE subject to translation-invariant Gaussian noise. This formula has a strong resemblance to that obtained for ordinary QFTs and allows us to extend the use of QFT tools in the analysis of the SPDE's effective potential. We furthermore demonstrate that much of the physical intuition regarding the effective action in QFTs also carries over into SPDEs. Finally, we offer a discussion of our results. A number of more technical issues are relegated to the Appendices. \section{Stochastic partial differential equations} \subsection{Elementary definitions} Consider the class of stochastic partial differential equations of the form \begin{equation} D \phi(\vec x,t) = F[\phi(\vec x,t)] + \eta(\vec x,t), \label{E:eom0} \end{equation} where $D$ is any linear differential operator, involving arbitrary time and space derivatives, which does {\em not} explicitly involve the field $\phi$. Typical examples are \begin{eqnarray} D &=& {\partial\over\partial t} - \nu \vec\nabla^2 \qquad \quad \hbox{Diffusion equation}, \\ D &=& {\partial^2\over\partial t^2} - \vec\nabla^2 \qquad \quad \hbox{Wave equation}, \\ D &=& {\partial\over\partial t} \qquad \qquad \qquad \hbox{Langevin equation}. \end{eqnarray} The function $F[\phi]$ is any forcing term, generally nonlinear in the field $\phi$. Typical examples are \begin{eqnarray} F[\phi] &=& +{\lambda\over2} (\vec \nabla\phi)^2 \qquad \hbox{in the Kardar--Parisi--Zhang (KPZ) equation}, \\ F[\phi] &=& P[\phi] \qquad\qquad\; \hbox{in reaction--diffusion--decay systems ($P$ is a polynomial)}, \\ F[\phi] &=& -{\delta H[\phi]\over\delta\phi} \qquad\;\; \hbox{in ``purely dissipative'' SPDEs}. \end{eqnarray} The forcing term will typically not contain any time derivatives, but this is not an essential part of the following analysis except insofar as time derivatives may complicate some of the Jacobian functional determinants that will be encountered below. Non-derivative terms linear in the field can be interpreted either as decay rates or (if a diffusion term is also present) as mass terms. They can be freely moved between the differential operator $D$ and the forcing term $F[\phi]$. If they are considered part of the forcing term then \begin{eqnarray} F[\phi] &=& -\gamma \phi \qquad \qquad \hbox{describes a decay term}; \\ F[\phi] &=& - \nu m^2 \phi \qquad \;\; \hbox{describes a mass term}. \end{eqnarray} The function $\eta(\vec x,t)$ is a random function of its arguments and describes the noise that we assume is present in the system. For the remainder of this paper, we consider field-independent additive noise. At this stage the nature and probability distribution of the noise are completely arbitrary and do not need to be specified. The noise represents our ignorance about precise details in the dynamics of the system. It could be due, for example, to fluctuations intrinsic to the dynamics (as in the case of Quantum Mechanics), or it could be thought of as representing the dynamics of short-scale degrees of freedom which have not completely decoupled from the macroscopic dynamics ({\em e.g.}, thermal or turbulent noise), or it could be a way of implementing ignorance of the {\em exact} initial or boundary conditions in the system. Noise can also be a way of summarizing the necessary truncation of the deterministic dynamics of a many body system when we try to describe it via a finite set of variables ({\em e.g.}, a truncated BBGKY hierarchy). If we think of turning off the noise, we do {\em not} require that the non-stochastic partial differential equation $D\phi = F[\phi]$ be derivable from an action principle ({\em i.e.}, the non-stochastic partial differential equation need not arise from a Lagrangian formalism). Nevertheless, once we include noise we demonstrate that the presence of noise automatically leads to a generalized action principle for the noisy system. It turns out that in the presence of Gaussian noise an equation of motion {\em proportional} to the factor $(D\phi - F[\phi])$ can always be derived by varying a well-defined ``classical'' action, and that the solutions to this equation of motion will coincide with those of the non-stochastic equation, provided a certain Jacobian determinant is nonsingular ({\em i.e.}, invertible). This is explained in full detail in Appendix B. \subsection{Some typical examples} An example of considerable interest is the reaction--diffusion--decay system where the SPDE is taken to be~\cite{vanKampen,Gardiner,HMPV-rdd} \begin{equation} \frac{\partial \phi}{\partial t} - \nu \vec\nabla^2 \phi = P[\phi] + \eta. \end{equation} This equation is used, for example, to describe the density $\phi(\vec x,t)$ of some chemical species as a function of space and time when the chemical is subject to both diffusion (via $\nu$) and reaction or decay [via $P(\phi)$, a polynomial in the density field]. Expanding out the first few terms, \begin{equation} P(\phi) = P_0 + P_1 \phi + P_2\phi^2 + P_3 \phi^3 +\cdots, \end{equation} we can identify $P_0$ with a constant (in space and time) source or sink, $- P_1$ with the decay rate, and $P_2$ with the reaction rate for the two-body reaction, {\em etc.} The noise accounts for random effects due to coupling to external sources, truncation of degrees of freedom, averaging over microscopic effects, {\em etc.} For $n$ species of chemical reactant the field $\phi$ is simply promoted to a vector in configuration space $\phi(\vec x, t) \rightarrow \phi_i(\vec x, t), \{i=1, \dots , n\}$. (The diffusion constant, $\nu$, and decay rates then become matrices, the noise a vector, and the polynomial $P_i(\phi_j)$ a vector-valued polynomial with tensorial coefficients.) A second well-known example is the massive KPZ equation (equivalent to the massive noisy Burgers equation) \cite{KPZ,MHKZ,Sun-Plischke,Frey-Tauber,HMPV-kpz} \begin{equation} \frac{\partial \phi}{\partial t} -\nu \vec\nabla^2 \phi = -\nu m^2 \phi +{\lambda\over2} (\vec\nabla\phi)^2 + \eta . \end{equation} In the fluid dynamical interpretation of the KPZ equation the fluid velocity is taken to be $\vec v = - \vec\nabla \phi$. This model problem leads to a form of ``turbulence'' which is known in the literature as Burgulence~\cite{Burgulence:1,Burgulence:2}. A third example is the enormous class of SPDEs known as ``purely dissipative'' systems~\cite{Zinn-Justin}. Purely dissipative systems have SPDEs of the form \begin{equation} {\partial \phi\over\partial t} = - {\delta H[\phi]\over\delta\phi(\vec x)} + \eta. \end{equation} These equations fall into our classification of general SPDEs as particular types of Langevin equations with $D=\partial_t$ and with a driving term that is a (functional) gradient $F[\phi] = -\delta H[\phi]/\delta\phi(\vec x)$. The nomenclature ``purely dissipative'' is justified by the fact that in the {\em absence} of noise these systems satisfy \begin{equation} {\partial H[\phi]\over\partial t} = -\int \left({\delta H[\phi]\over\delta\phi(\vec x)}\right)^2 d^d \vec x \leq 0. \end{equation} Note that the reaction--diffusion--decay system can be interpreted as an example of a purely dissipative system if we take $D=\partial_t$ and \begin{equation} H_{ RDD}[\phi] = \int \left[ {\nu\over2}(\vec \nabla\phi)^2 + \int_0^{\phi(x)} P(\tilde\phi) \; {\mathrm{d}}\tilde\phi \right] {\mathrm{d}} \vec x. \end{equation} On the other hand, the KPZ system is {\em not} a purely dissipative system, \begin{eqnarray} {\delta F_{ KPZ}[\phi(x)] \over \delta\phi(y)} &=& \nu m^2 \delta(x-y) + \lambda \vec \nabla_x \phi(x) \cdot \vec \nabla_x \delta(x-y) \nonumber\\ &=& \nu m^2 \delta(x-y) - \lambda \vec \nabla_y \phi(y) \cdot \vec \nabla_y \delta(x-y) \nonumber\\ &\neq& {\delta F_{ KPZ}[\phi(y)] \over \delta\phi(x)}. \end{eqnarray} The class of purely dissipative SPDEs is a very wide one, but there are many SPDEs that are not of purely dissipative type. We do {\em not} want to restrict attention to purely dissipative systems in this article and keep the discussion as general as possible. \section{Stochastic averages, characteristic functional, Feynman rules} We will focus on the stochastic partial differential equation \begin{equation} D\phi(\vec x,t) = F[\phi(\vec x,t)] + \eta(\vec x,t), \label{E:stochastic1} \end{equation} and analyze it using functional integral techniques: Feynman diagrams, the effective action, and the effective potential. We develop the field theory via the most direct route, with no conjugate fields present. We postpone to subsequent papers more technically involved approaches such as the Martin--Siggia--Rose Lagrangian (with its extra unphysical conjugate fields used for book-keeping purposes) and the hidden BRST supersymmetry implicit in these stochastic differential equations~\cite{MSR,De-Dominicis-Peliti,Zinn-Justin,Gozzi}. In this section we develop the necessary tools to construct the basic field theory and non--equilibrium statistical mechanics associated with equation (\ref{E:stochastic1}). We will assume {\em uniqueness} of the solution to (\ref{E:stochastic1}), and in order to calculate the characteristic functional, we will introduce an {\em ensemble average} over noise realizations, and the notion of {\em delta functionals}. Once the {\em characteristic functional} is available, we find it useful to introduce {\em ghosts \'a la Faddeev--Popov} before deriving the {\em Feynman rules}. We will only need to make one assumption about the noise: that it be Gaussian, {\em i.e.,} that all its cumulants are vanishing except the first, $\langle \eta(\vec x,t)\rangle$, and second, $G_{\eta} = \langle \eta(\vec x,t)\eta(\vec x',t')\rangle$. (See {\em e.g.},~\cite{Marcienkiewicz}.) \subsection{Step 1: Uniqueness} Let us assume that the partial differential equation (\ref{E:stochastic1}), plus initial conditions, is a well-posed problem. Thus, given a particular realization of the noise, $\eta$, the differential equation is assumed to have a unique solution which we designate as \begin{equation} \phi_{ soln}(\vec x,t|\eta). \end{equation} This assumption is relatively mild but does imply that the nonlinearity is sufficiently weak so as not to drive us past a bifurcation point. On the other hand, it is known that noise in concert with non--linearities can lead to the phenomenon of delayed bifurcation in non--linear parabolic SPDEs \cite{Lythe}. If the partial differential equation is ill-posed, in the sense that the solutions are not unique, additional analysis must be developed on a case-by-case basis. A specific example of this behavior is spontaneous symmetry breaking in QFT, which causes the naive loop expansion to violate the convexity properties of the effective potential. This situation must be dealt with by an improved loop expansion~\cite{Fujimoto,Haymaker,Bender-Cooper}. \subsection{Step 2: Ensemble average} For any function $Q(\phi)$ of the field $\phi$ we introduce the ensemble average (over the noise), defined by \begin{equation} \langle Q(\phi) \rangle \equiv \int ({\cal D} \eta) \; {\cal P} [\eta] \; Q(\phi_{ soln}(\vec x, t|\eta)), \label{E:average1} \end{equation} where $ {\cal P} [\eta]$ is the probability density functional of the noise. It is normalized to $1$, but is otherwise completely arbitrary, that is \begin{equation} \int ({\cal D} \eta) \; {\cal P} [\eta] = 1. \end{equation} The symbol ${\cal D}\eta$ indicates a functional integral over all instances (or realizations) of the noise. \subsection{Step 3: Delta functionals} We next use a functional delta function to write the following identity \begin{eqnarray} \phi_{ soln}(\vec x,t|\eta) &\equiv& \int ({\cal D} \phi) \; \phi \; \delta[\phi - \phi_{ soln}(\vec x,t|\eta)] \nonumber\\ &=& \int ({\cal D} \phi) \; \phi \; \delta[D \phi- F[\phi] - \eta] \; \sqrt{ {\cal J} {\cal J}^\dagger }, \end{eqnarray} where we have performed a change of variables and introduced the Jacobian functional determinant, defined by \begin{equation} {\cal J} \equiv \det\left( D - {\delta F\over\delta\phi} \right), \end{equation} and its adjoint \begin{equation} {\cal J}^\dagger \equiv \det\left( D^\dagger - {\delta F\over\delta\phi}^\dagger \right). \end{equation} The above is just the functional analogue of a standard delta function result: If $f(x) =0$ has a unique solution at $x=x_0$, then \begin{equation} x_0 = \int {\mathrm{d}} x \; x \; \delta(x-x_0) = \int {\mathrm{d}} x \; x \; \delta(f(x)) \; |f'(x)| = \int {\mathrm{d}} x \; x \; \delta(f(x)) \; \sqrt{f'(x) [f'(x)]^*}. \end{equation} The delta function forces one to pick up only one contribution from the solution of the equation $f(x)=0$, and the derivative is there to provide the correct measure to the integral. In the functional case the derivative becomes a determinant. It is in fact the Jacobian determinant associated with the change of variables from $\phi$ to $D\phi-F[\phi]$. It is now easy to see that one also has the identity \begin{equation} Q(\phi_{ soln}(\vec x,t|\eta)) \equiv \int ({\cal D} \phi) \; Q(\phi) \; \delta\left( D \phi- F[\phi] - \eta\right) \; \sqrt{{\cal J} {\cal J}^\dagger }. \end{equation} Furthermore, the ensemble average over the noise, equation (\ref{E:average1}), becomes \begin{equation} \langle Q(\phi) \rangle = \int ({\cal D} \eta) \; ({\cal D} \phi) \; {\cal P} [\eta] \; Q(\phi) \; \delta\left( D \phi- F[\phi] - \eta\right) \; \sqrt{{\cal J} {\cal J}^\dagger }. \end{equation} The noise integral is easy to perform, with the result that for arbitrary stochastic averages one has \begin{equation} \langle Q(\phi) \rangle = \int ({\cal D} \phi)\; {\cal P} [D\phi-F[\phi]] \; Q(\phi) \; \sqrt{{\cal J} {\cal J}^\dagger }. \end{equation} We see from this equation that the effect of the noise only appears in the stochastic average through its probability distribution $ {\cal P} [D\phi-F[\phi]]$. It is worthwhile to point out that the main difference, at this stage of the formalism, between the present ``minimal'' approach and that of MSR lies in the way the delta functional is handled. In MSR, instead of integrating directly over the noise, as is done here, the delta functional is replaced by its functional Fourier integral representation. This is the step wherein the conjugate field enters. If this latter route is taken, the noise integration can be performed exactly only for Gaussian noise. In the minimal formalism by contrast, the integration over the noise can be done exactly for arbitrary noise. It thus lends itself immediately for handling non-Gaussian systems: For general noise distributions we can explicitly write down the probability distribution for the fields as \begin{equation} {\bf P} [\phi] = {\cal P} [D\phi-F[\phi]] \; \sqrt{{\cal J} {\cal J}^\dagger }. \end{equation} We will not further explore the possibility of arbitrary noise in this paper, since Gaussian noise (which manifestly does {\em not} imply Gaussian fluctuations of the fields) is already sufficiently general to be of great practical interest. The presence of the functional determinant is essential: it must be kept to ensure proper counting of the solutions to the original stochastic differential equation. In QFT this functional determinant is known as the Faddeev--Popov determinant and is essential in maintaining unitarity~\cite{Weinberg-1,Rivers}, {\em i.e.}, conservation of probability. In some particular cases the functional determinant is field-independent, and it is safe to neglect it. We discuss this more fully in Appendix A and in the companion papers~\cite{HMPV-rdd,HMPV-kpz}, but for the sake of generality we will carry these determinants along (with little extra cost) for the rest of this paper. \subsection{Step 4: Characteristic functional (Partition function)} A particularly useful quantity is the generating functional, or characteristic functional (partition function), defined by taking \begin{equation} Q(\phi) = \exp\left( \int {\mathrm{d}}^d\vec x \; {\mathrm{d}} t \; J(\vec x,t) \; \phi(\vec x,t) \; \right) \end{equation} in equation (\ref{E:average1}). We define it as follows \begin{eqnarray} Z[J] &\stackrel{ def}{=}& \left\langle \exp \left( \int {\mathrm{d}} x \; J(x) \; \phi(x) \right) \right\rangle \\ &=& \int ({\cal D} \phi)\; {\bf P} [\phi] \exp\left( \int {\mathrm{d}} x \; J \; \phi \; \right) \; \\ &=& \int ({\cal D} \phi)\; {\cal P} [D \phi- F[\phi]] \; \exp\left( \int {\mathrm{d}} x \; J \; \phi \; \right) \; \sqrt{{\cal J} {\cal J}^\dagger }, \label{E:characteristic0} \end{eqnarray} with an obvious condensation of notation, ${\mathrm{d}} x = {\mathrm{d}}^d\vec x \; {\mathrm{d}} t$. When there is no risk of confusion we will suppress the ${\mathrm{d}} x$ completely. This key result will enable us to calculate the effective action and the effective potential in a direct way. \subsection{Step 5: Gaussian noise} We will now make some assumptions about the noise: We assume it to be Gaussian. Without loss of generality we can take the noise to have zero mean, since if the mean is nonzero we can always redefine the forcing term $F[\phi]$ to make the noise have zero mean. We therefore take the noise to be Gaussian of zero mean, so that the only non-zero cumulant is the second order one. We do not need to make any more specific assumptions about the functional realization of the noise: the noise might (for instance) be white, power-law, colored, pink, $1/f$-noise, or shot noise, and our considerations below apply to all of these cases. As long as the noise is Gaussian its probability distribution can be written as \begin{equation} {\cal P} [\eta] = {1\over\sqrt{\det(2\pi G_\eta)}} \exp\left( - {\scriptstyle{1\over2}} \int \int {\mathrm{d}} x \; {\mathrm{d}} y \; \eta(x) \; G_\eta^{-1}(x,y) \; \eta(y) \right). \end{equation} The characteristic functional (partition function) is thus seen from equation (\ref{E:characteristic0}) to be \begin{eqnarray} Z[J] &=& {1\over\sqrt{\det(2\pi G_\eta)}} \int ({\cal D} \phi)\; \sqrt{{\cal J} {\cal J}^\dagger } \exp\left( \int J \phi \right) \; \exp\left( - {\scriptstyle{1\over2}} \int \int (D\phi-F[\phi]) G_\eta^{-1} (D\phi-F[\phi]) \right). \label{E:characteristic1} \end{eqnarray} This characteristic functional (partition function) contains all the physics of the model since it allows for the calculation of averages, correlation functions, thermodynamic variables, {\em etc.} Note that the noise has been completely eliminated and survives only through the explicit appearance of its two-point correlation function in the above. Since the characteristic functional is now given as a path integral over the physical field, all the standard machinery of statistical field theory (and quantum field theory) can be brought to bear~\cite{Textbooks}. See, for example~\cite{Zinn-Justin,Weinberg-1,Weinberg-2,Rivers}, and~\cite{Bjorken-Drell-1}---\cite{Le-Bellac}. This formula for the partition function demonstrates that (modulo Jacobian determinants) all of the physics of any stochastic differential equation can be extracted from a functional integral based on the ``classical action'' \begin{equation} {\cal S}_{ classical} = \frac{1}{2} \int \int (D\phi-F[\phi]) G_\eta^{-1} (D\phi-F[\phi]) \; . \end{equation} This ``classical action'' is a generalization of the Onsager--Machlup action~\cite{Onsager-Machlup}. The Onsager--Machlup paper dealt with stochastic differential equations rather than partial differential equations (mechanics rather than field theory), and was limited to noise that was temporally white. As their formalism was developed with the notions of linear response theory in mind, Onsager and Machlup assumed the ``forcing term'' $F[\phi]$ to be linear, so that {\em both} the noise and the field fluctuations were Gaussian. In our formalism all these assumptions can be relaxed: the forcing term can be nonlinear and in general the field fluctuations will not be Gaussian even if the noise is Gaussian. \subsection{Step 6: Faddeev--Popov ghosts} We mentioned previously that the Jacobian determinant is often (not always) field-independent. This is a consequence of the causal structure of the theory as embodied in the fact that we are only interested in {\em retarded} Green functions. The situation here is in marked contrast with that in QFT where the relativistic nature of the theory forces the use of {\em Feynman} Green functions ($+i\epsilon$ prescription). As we explain in Appendix A, this change radically alters the behavior of the functional determinant. In order to avoid too many special cases, and to have a formalism that can handle both constant and field-dependent Jacobian factors, we exponentiate the determinant via the introduction of a pair of Faddeev--Popov ghost fields~\cite{Zinn-Justin,Weinberg-1}, \begin{equation} \label{E:ghost} {\cal J} \equiv \det\left( D - {\delta F\over\delta\phi} \right) = {1\over\det(2\pi I)} \int ({\cal D} [g^\dagger,g]\,) \; \exp\left( - {\scriptstyle{1\over2}} \int g^\dagger \left[ D - {\delta F\over\delta\phi} \right] g \right), \end{equation} where $I$ is the identity operator on spacetime. The $g$ field is a so-called (complex) scalar ghost field. It is a field of anti-commuting complex variables and behaves in a manner similar to an ordinary scalar field except that there is an extra minus sign for each ghost loop. We also need to use the {\em conjugate} ghost field $g^\dagger$ to handle the determinant. We should point out that if the operator $D - (\delta F/\delta\phi)$ is self-adjoint then $\cal J = {\cal J}^\dagger$. In this case $\sqrt{ {\cal J} {\cal J}^\dagger}$ reduces to $|{\cal J}| = |{\cal J}^\dagger|$. In QFTs the relevant operators occurring in the Jacobian determinants are obtained from second functional derivatives of the action and are automatically self-adjoint. In contrast, for SPDEs there is no guarantee that $D - (\delta F/\delta\phi)$ be self-adjoint, and in fact for the examples previously discussed (KPZ, reaction--diffusion--decay, and purely dissipative) this operator is not self adjoint. Instead we rely on the much weaker property that the operator $D - (\delta F/\delta\phi)$ is {\em real} in order to write $\sqrt{ {\cal J} {\cal J}^\dagger} = \sqrt{{\cal J} {\cal J}^*} = |{\cal J}|$. In all cases we are interested in the relevant operators are not only real, but positive, so that the absolute value symbol can be ignored and the characteristic functional equation (\ref{E:characteristic1}) is given by \begin{eqnarray} Z[J] &=& {1\over\sqrt{\det[(2\pi)^3 G_\eta]}} \int ({\cal D} \phi)\; ({\cal D} g) \; ({\cal D} g^\dagger) \; \exp\left( - {\scriptstyle{1\over2}} \int \int (D\phi-F[\phi]) G_\eta^{-1} (D\phi-F[\phi]) \right) \; \nonumber\\ &&\qquad \exp\left( - {\scriptstyle{1\over2}} \int g^\dagger \left[ D - {\delta F\over\delta\phi} \right] g \right) \; \exp\left( \int J \phi \right). \label{E:characteristic2} \end{eqnarray} This procedure trades off the functional determinants for two extra functional integrals. The advantage of this procedure becomes clear when one develops the perturbation theory. (This Faddeev--Popov trick for exponentiating the Jacobian determinant is also essential in finding the hidden BRST supersymmetry.) It must be noted that neither pair of ghost field variables couple to an external source. This means they can only appear in internal lines in Feynman diagrams, a fact that will be used later on when we discuss loops and loop-counting. In Appendix A we take a closer look at the Jacobian functional determinant, its causal structure, and its specific form for local driving forces. In the latter part of this Appendix we make use of the perturbation theory based on Feynman diagrams to evaluate this functional determinant from another perspective. \subsection{Step 7: Feynman rules} With the partition function in the form given above, (with two independent ghosts), it is now easy to develop a formal Feynman diagram expansion. We wish to treat the driving term $F[\phi]$ as the perturbation and expand around the free-field theory defined by setting $F=0$. With this convention the free action is, explicitly, \begin{equation} {\cal S}_{ free} \equiv \int \int \left\{ {\scriptstyle{1\over2}} \; [D\phi] \; G_\eta^{-1} \; [D\phi] \right\} \; {\mathrm{d}} x \; {\mathrm{d}} y + \int \left\{ {\scriptstyle{1\over2}} \; g^\dagger \; D \; g \right\} dx. \end{equation} There are two particle propagators in this free action, one for the $\phi$ field, and two for the ghost fields. Formally \begin{eqnarray} G_{\phi\phi} &=& [ D^\dagger \; G_\eta^{-1} \; D ]^{-1} = [ D^{-1} \; G_\eta (D^\dagger)^{-1} ]; \\ G_{g^\dagger g} &=& [ D ]^{-1}. \end{eqnarray} Here $D^\dagger$ is the adjoint operator of $D$, defined by partial integration. (For instance if $D=\partial_t -\nu \vec \nabla^2$ then $D^\dagger = -\partial_t - \nu \vec \nabla^2$.) In the interests of generality, we reiterate the fact that we have not assumed translation invariance for the noise, (although the noise is now Gaussian.) The momentum-frequency representation for the propagators is \begin{eqnarray} G_{\phi\phi}(\vec k_1,\omega_1;\vec k_2,\omega_2) &=& { G_\eta(\vec k_1,\omega_1;\vec k_2,\omega_2) \over D^\dagger(\vec k_1,\omega_1) \; D(\vec k_2,\omega_2)}; \\ G_{g^\dagger g}(\vec k,\omega) &=& {1\over D(\vec k,\omega)}; \end{eqnarray} The Feynman vertices come from the interaction piece of the action, which in this convention is \begin{equation} {\cal S}_{ interaction} = \int \int \left\{ -[D\phi] \; G_\eta^{-1} \; F[\phi] + {\scriptstyle{1\over2}} F[\phi] \; G_\eta^{-1} \;F[\phi] \right\} {\mathrm{d}} x \,{\mathrm{d}} y - \int \left\{ {\scriptstyle{1\over2}} \; g^\dagger \; {\delta F\over\delta\phi} \; g\right\} {\mathrm{d}} x. \end{equation} The nature of the vertices (obtained by functional differentiation with respect to the fields present in the theory) depends on the structure of the forcing term $F[\phi]$. In the meantime we formally assert \begin{eqnarray} &\phi-F[\phi]\hbox{ vertex:}& \qquad -{D(\vec k_1,\omega_1) \; \phi \; F[\phi]\over G_\eta(\vec k_1,\omega_1;\vec k_2,\omega_2)}; \\ &F[\phi]-F[\phi] \hbox{ vertex:}& \qquad + {\scriptstyle{1\over2}} \; { F[\phi] \; F[\phi] \over G_\eta(\vec k_1,\omega_1;\vec k_2,\omega_2)}; \\ &\hbox{ghost vertex:}& \qquad - {\scriptstyle{1\over2}} \; g^\dagger\; {\delta F[\phi]\over\delta\phi}\; g. \end{eqnarray} Note that to turn these schematic Feynman rules into practical computational tools we will need to assume that $F[\phi]$ is some specific local functional of the field $\phi$. (Typically a polynomial or polynomial with derivatives.) When we define the effective action we will again see that the formalism can be successfully developed even for non-translation-invariant noise, and this derivation of the Feynman rules matches the generality of the definition of the effective action. This concludes, for now, the most general aspects of the discussion. When it comes to actual calculations in specific models, the majority of these models have noise that is not only Gaussian but is also translation invariant. In the interests of simplicity, we now (finally, and only for the rest of this particular section) indicate the effects of assuming translation invariance for the noise. This lets us take simple Fourier transforms in the difference variable $x-y$ (more precisely: $\vec x-\vec y$ and $t_x-t_y$) to see that in momentum-frequency space \begin{eqnarray} G_{\phi\phi}(\vec k,\omega) &=& { G_\eta(\vec k,\omega)\over D^\dagger(\vec k,\omega) \; D(\vec k,\omega)} = { G_\eta(\vec k,\omega)\over D(-\vec k,-\omega) \; D(\vec k,\omega)}; \\ G_{g^\dagger g}(\vec k,\omega) &=& {1\over D(\vec k,\omega)}. \end{eqnarray} The Feynman diagram vertices are now ($\phi$ and $g$ are here understood to be Fourier transformed) \begin{eqnarray} &\phi-F[\phi]\hbox{ vertex:}& \qquad -{D(\vec k,\omega) \;\phi \; F[\phi]\over G_\eta(\vec k,\omega)}; \\ &F[\phi]-F[\phi] \hbox{ vertex:}& \qquad + {\scriptstyle{1\over2}} { F[\phi] \; F[\phi] \over G_\eta(\vec k,\omega)}; \\ &\hbox{ghost vertex:}& \qquad - {\scriptstyle{1\over2}} \; g^\dagger \; {\delta F[\phi]\over\delta\phi} \; g. \end{eqnarray} As always, there is a certain amount of freedom in writing down the Feynman rules. It is always possible to take part of the quadratic piece in the total action and move it from the free action to the interaction term or vice versa. We have already seen that a linear term ({\em e.g.}, $\nu m^2 \phi$) in the forcing function $F[\phi]$ can with equal facility be reassigned to the differential operator $D$ via the scheme $D \to D - \nu m^2$, $F[\phi]\to F[\phi] - \nu m^2 \phi$. This procedure can always be used to completely eliminate any linear term in $F[\phi]$. Similar but more complicated behaviour occurs if the forcing function contains both constant and quadratic pieces ({\em e.g.}, $a+b\phi^2$). With the conventions given above the interaction term contains (at least) a ``cosmological constant'', $ {\scriptstyle{1\over2}} a^2 \int \int G_\eta^{-1}(x,y) dx dy$, a quadratic piece, $ab \int \int G_\eta^{-1}(x,y) \phi^2(y) dx dy$, and a $\phi^4$ interaction. The quadratic piece could be moved into the free action at the cost of making the expression for the free propagator a little more complicated. This freedom in writing down the Feynman rules does not imply any ambiguity in the physical results: Moving quadratic pieces around from the interaction term to the free action will modify the Feynman rules but will not affect any physical quantities. \section{Effective Action: Loop expansion} In order to set up the formalism for the effective action, and its loop expansion, it is useful to first separate the two-point function for the noise into a {\em shape}, $g_2(x,y)$, and a constant {\em amplitude}, $ {\cal A} $, via the correspondence \begin{equation} G_\eta(x,y) \mathop{\stackrel{ def}{=}} {\cal A} \; g_2(x,y). \end{equation} For the case of Gaussian white noise this is automatically satisfied by definition: $G_\eta(x,y) \to {\cal A} \; \delta(x-y)$. For more general Gaussian noises (which describe, for instance, the effects of small scale degrees of freedom not fully decoupled from the physics of $\phi$, such as in the case of a heat bath into which $\phi$ has been immersed) this form of the two-point function allows the interpretation of the noise intensity $ {\cal A} $ as a characterization of the bath-system coupling. This will become more evident when we compare the ``effective potential'' in noisy environments, equation (\ref{E:general}) below, with the same object for zero temperature quantum field theory, equation (\ref{E:general1}). The normalization of the shape function $g_2(x,y)$ is essentially arbitrary, and any convenient normalization will suffice. Another advantage of singling out the intensity parameter $ {\cal A} $ is that it is the loop-counting parameter for this formulation of SPDEs. To see this, one starts by writing the characteristic functional (with external sources rescaled for convenience) as \begin{eqnarray} Z[J] &=& {1\over\sqrt{\det[(2\pi)^3 G_\eta]}} \int ({\cal D} \phi)\; ({\cal D} g) \; ({\cal D} g^\dagger) \; \exp\left( - {\scriptstyle{1\over2}} \int \int { (D\phi-F[\phi]) g_2^{-1} (D\phi-F[\phi]) \over {\cal A} } \right) \; \nonumber\\ &&\qquad \exp\left( - {\scriptstyle{1\over2}} \int g^\dagger \left[ D - {\delta F\over\delta\phi} \right] g \right) \; \exp\left( {\int J \; \phi\over {\cal A} } \right). \label{E:characteristic3} \end{eqnarray} The generating function (Helmholtz free energy in statistical field theory) for connected correlation functions is defined by \begin{equation} W[J] = + {\cal A} \left\{ \ln Z[J] - \ln Z[0] \right\}. \end{equation} The effective action (Gibbs free energy in statistical field theory) is then defined {\em non-perturbatively} in terms of $W[J]$ by taking its Legendre transform~\cite{Weinberg-2,Non-perturbative} \begin{equation} \Gamma[\phi;\phi_0] = - W[J] + \int \phi \; J; \qquad\qquad {\delta W[J]\over \delta J} = \phi; \qquad\qquad {\delta \Gamma[\phi;\phi_0]\over \delta \phi} = J. \end{equation} Here $\phi_0$ is some suitable background (mean) field, which is taken to be the stochastic expectation value of $\phi$ in the absence of external sources, $J=0$. It is often but not always zero, and we retain it for generality. The previous equation defines the non-perturbative effective action. For any specific example the previous equation is not very useful, and we often have to restrict ourselves to a perturbative calculation of the effective action, after singling out an expansion parameter. One can always develop a Feynman diagram expansion provided that the classical action can be separated into a ``quadratic piece'' and an ``interacting term'', as we have already done. In the loop expansion the sum of all connected diagrams coupled to external sources $J(x)$ is exactly $W[J]$ as defined above, and the effective action $\Gamma[\phi;\phi_0]$ corresponds to all (amputated) one-particle irreducible graphs (1PI), that is, Feynman diagrams that cannot be made disconnected by cutting only one propagator. In the following argument, we will be considering diagrams contributing to the effective action. Recall that ghosts can only appear as internal lines, since they are not coupled to external sources. To see the role of the amplitude $ {\cal A} $ as loop-counting parameter, note that each field propagator is proportional to $ {\cal A} $ while each ghost propagator is independent of $ {\cal A} $. The vertices that do not include ghosts are proportional to $ {\cal A} ^{-1}$, while ghost vertices are independent of $ {\cal A} $. Thus each Feynman diagram contributing to the effective action is proportional to $A^{I_\phi-V_\phi}$, where $I_\phi$ is the number of non-ghost propagators, and $V_\phi$ is the number of non-ghost vertices. But each ghost vertex is attached to exactly two ghost propagators (except for tadpole ghost loops), and each ghost propagator is attached to exactly two ghost vertices (except for tadpole ghost loops). In the case of tadpole ghost loops, exactly one propagator is attached to exactly one ghost vertex. This implies that if one assigns a factor $ {\cal A} $ to each ghost propagator and a factor of $ {\cal A} ^{-1}$ to each ghost loop then one will not change the total number of factors of $ {\cal A} $ assigned to the Feynman diagram. Thus the Feynman diagrams are proportional to $ {\cal A} ^{I-V}$ where $I$ is the total number of (internal) propagators in the Feynman diagram and $V$ is the total number of vertices, now including ghosts. It is the result of a standard topological theorem that for any graph (not just any Feynman diagram) $ I - V = L - 1 $, where $L$ is the number of loops~\cite{Weinberg-1,Rivers,Itzykson-Zuber}. It is then easy to see that field theories based on SPDEs exhibit exactly the same loop-counting properties as QFTs except that the loop-counting parameter is now the {\em amplitude} of the noise two-point function (instead of Planck's constant $\hbar$). The only subtle part of the argument has been in dealing with the Faddeev--Popov ghosts, and it is important to realize that this argument is completely independent of the details of the differential operator $D$ and the forcing term $F[\phi]$. When it comes to calculating the diagrams contributing to the effective action, the extra explicit factor of $ {\cal A} $ inserted in the definition of $W[J]$ above guarantees that the 1PI graphs contribute to $\Gamma[\phi;\phi_0]$ with a weight that is exactly $ {\cal A} ^L$. This demonstrates that $ {\cal A} $ is a {\em bona fide} expansion parameter. At this point, it becomes natural to make a comparison with the MSR (Martin--Siggia--Rose) formalism for the calculation of the effective action in Stochastic Field Theories, where one introduces a field {\em conjugate} to $\phi$. Historically, this conjugate field first arose in setting up a variational approach to the diffusion equation, ({\em cf.} Morse and Feshbach {\em loc. cit.}). The following remarks will help one to understand the differences and the complementarity of our approach to the MSR approach; the bottom line is related to technical issues associated to proving all-orders renormalizability~\cite{MSR,De-Dominicis-Peliti,Zinn-Justin}. The direct approach developed in this paper is akin to the ghost-free axial gauge of QCD or the so-called unitary gauge in the standard model of particle physics: This is a formalism well-adapted to isolating the physical degrees of freedom, at least perturbatively, but is not well-adapted to proving the all-orders renormalizability of the theory. (Proving one-loop renormalizability for specific theories is not too difficult, and we will address this issue in a pair of companion papers~\cite{HMPV-rdd,HMPV-kpz}.) In analogy with the situation in QFT, one has three possible responses to this state of affairs: \begin{enumerate} \item Use the MSR formalism for all calculations. This is comparable to using BRST--invariant versions of the standard model of particle physics to calculate scattering cross-sections and decay rates. (That is, overkill.) \item One could appeal to the fact that the SPDEs considered in this paper are hardly likely to be thought of as fundamental theories in the particle physics sense; These SPDEs are much more like ``effective field theories'', in that the noise and fluctuations in real physical systems are manifestations of our lack of knowledge of the short-distance physics. Viewed as effective theories, renormalizability is no longer the main guiding light it was once thought to be~\cite{Weinberg-2}. \item At a very {\em practical} level one can choose to be guided by experience with quantum field theories. It is well known that one-loop physics is often sufficient for extracting most of the physical information from a system. Calculations beyond one-loop, while certainly important at a fundamental level, are often more than is really needed. One of the great technical simplifications of one-loop physics is that, via zeta function technology, essentially any field theory can be regularized at one-loop without excessive complications~\cite{Dowker,DeWitt,BVW}. \end{enumerate} For these reasons we will now restrict our attention to a one-loop calculation (apart from the discussion of Feynman diagrams and the loop expansion everything up to this point has been valid non-perturbatively, while those discussions were still valid to all orders in perturbation theory). In the next section we calculate the one-loop effective action. \section{Effective Action: One Loop} It is well known that the effective action for a field theory can be obtained by performing a Legendre transform on the logarithm of the characteristic functional (partition function). Writing \begin{equation} Z[J] = \int {\cal D}\phi \exp\left({ - {\cal S}[\phi] + \int J\; \phi \over a}\right), \label{E:characteristic4} \end{equation} where {\em a} is the parameter characterizing the fluctuations, one gets for the one-loop effective action (first order in {\em a}) \begin{equation} \Gamma[\phi;\phi_0] = {\cal S}[\phi] - {\cal S}[\phi_0] + {\scriptstyle{1\over2}} a \left\{ \ln \det ({\cal S}_2[\phi]) - \ln \det ({\cal S}_2[\phi_0]) \right\} + O(a^2). \label{E:effective} \end{equation} Here ${\cal S}_2 \mathop{\stackrel{ def}{=}} \delta^2 {\cal S}/\delta \phi(x) \delta \phi(y)$ is the matrix of second order functional derivatives of the action ${\cal S}[\phi]$ (often called the Jacobi field operator). For QFT the loop-counting parameter $a$ is Planck's constant $\hbar$, and ${\cal S}_2$ is a second-order partial differential operator that depends on the field $\phi$ via some potential-like term. The determinants of partial differential operators can be defined and calculated by a variety of techniques. The notation ${\cal S}[\phi_0]$ is actually shorthand for ${\cal S}[\langle\phi[J=0]\rangle]$, and for a symmetric ground state ($\langle\phi[J=0]\rangle=0$) one often has ${\cal S}[0]=0$. These terms contribute a constant offset to the effective action. In QFT these terms are interpreted as a field-independent contribution to the vacuum energy and are traditionally ignored, although in the context of cosmology, they contribute (sometimes catastrophically) to the cosmological constant. In the interest of generality we will make them explicit. When we consider field theories based on SPDEs, the loop-counting parameter {\em a} becomes $ {\cal A} $, which we singled out as the amplitude for the noise, and the bare action in equation (\ref{E:characteristic3}) is replaced by equations (\ref{E:characteristic1}) and (\ref{E:characteristic2}) \begin{eqnarray} {\cal S}[\phi] &\to& {\scriptstyle{1\over2}} \int \int \left\{ (D\phi-F[\phi]) g_2^{-1} (D\phi-F[\phi]) \right\} {\mathrm{d}}^d \vec x \; {\mathrm{d}} t\; {\mathrm{d}}^d \vec y \; {\mathrm{d}} t' - {\scriptstyle{1\over2}} {\cal A} \; \left( \ln {\cal J} + \ln {\cal J}^\dagger \right) \\ &=& {\cal S}_{ classical}[\phi] - {\scriptstyle{1\over2}} {\cal A} \; \left( \ln {\cal J} + \ln {\cal J}^\dagger \right), \label{E:bareaction} \end{eqnarray} where on the second line we have denoted by ${\cal S}_{ classical}[\phi]$ the double integral in the previous line. This is the quantity that we have previously defined as the nonlinear generalization of the Onsager--Machlup action to arbitrary Gaussian noise~\cite{Onsager-Machlup}. The noise at this stage is Gaussian, and does not need to be translation invariant. We have explicitly kept the Jacobian functional determinant. Inserting equation (\ref{E:bareaction}) into the formula for the one-loop effective action [equation (\ref{E:effective})], we obtain the following general result (applicable to any SPDE) \begin{eqnarray} \Gamma[\phi;\phi_0] &=& {\cal S}_{ classical}[\phi] - {\cal S}_{ classical}[\phi_0] \nonumber\\ &&\qquad + {\cal A} \left\{ {\scriptstyle{1\over2}} \ln \det ({\cal S}_2[\phi]) - {\scriptstyle{1\over2}} \ln \det ({\cal S}_2[\phi_0]) - {\scriptstyle{1\over2}} \ln {\cal J}[\phi]- {\scriptstyle{1\over2}} \ln {\cal J}^\dagger[\phi] + {\scriptstyle{1\over2}} \ln {\cal J}[\phi_0] + {\scriptstyle{1\over2}} \ln {\cal J}^\dagger[\phi_0] \right\} \nonumber\\ &&\qquad + O({\cal A}^2). \end{eqnarray} To make this more explicit, the fluctuation operator ${\cal S}_2(\phi)$ ({\em aka} Jacobi field operator) is \begin{equation} {\cal S}_2[\phi] = \left( D^{\leftarrow} - {\delta F\over \delta\phi}^{\leftarrow} \right) g_2^{-1} \left( D - {\delta F\over \delta\phi} \right) - (D\phi-F[\phi]) g_2^{-1} {\delta^2 F\over\delta\phi\;\delta\phi}. \end{equation} Here the $\leftarrow$ indicates that these operators should be thought of as acting to the left. Also $g_2^{-1}(x,y)$ is to be understood as a ``matrix'' with implicit sums over the indices $x,y$ ({\em i.e.}, integrations over the variables.) Note that if $F[\phi]$ contains derivatives of $\phi$ then $\delta F/\delta\phi$ will be a differential operator. Performing an integration by parts, this can be converted to a statement about the adjoint operator acting to the right, {\em i.e.,} we can re-write ${\cal S}_2[\phi]$ as \begin{equation} {\cal S}_2[\phi] = \left( D^\dagger - {\delta F\over \delta\phi}^\dagger \right) g_2^{-1} \left( D - {\delta F\over \delta\phi} \right) - (D \phi-F[\phi]) g_2^{-1} {\delta^2 F\over\delta\phi\;\delta\phi}. \end{equation} Putting all this together gives the following one-loop result for the effective action \begin{eqnarray} \label{E:Gamma-1} \Gamma[\phi;\phi_0] &=& {\scriptstyle{1\over2}} \int \int {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; {\mathrm{d}}^d \vec y \; {\mathrm{d}} t' \left\{ (D \phi-F[\phi]) g_2^{-1} (D \phi-F[\phi]) \right\} - {\scriptstyle{1\over2}} {\cal A} \; \left( \ln {\cal J} + \ln {\cal J}^\dagger \right) \nonumber\\ &&+ {\scriptstyle{1\over2}} {\cal A} \ln \det \left[ \left( D^\dagger - {\delta F\over \delta\phi}^\dagger \right) g_2^{-1} \left( D - {\delta F\over \delta\phi} \right) - (D \phi-F[\phi]) g_2^{-1} {\delta^2 F\over\delta\phi\;\delta\phi} \right] \nonumber\\ &&- \left( \phi \to \phi_0 \right) + O( {\cal A} ^2). \end{eqnarray} Grouping together the terms proportional to $ {\cal A} $, and using the representation of the functional determinant, enables us to rewrite the above in the alternative form \begin{eqnarray} \label{E:Gamma-2} \Gamma[\phi;\phi_0] &=& {\scriptstyle{1\over2}} \int \int {\mathrm{d}}^d \vec x\; {\mathrm{d}} t \; {\mathrm{d}}^d \vec y \; {\mathrm{d}} t' \left\{ (D \phi-F[\phi]) g_2^{-1} (D \phi-F[\phi]) \right\} \nonumber\\ &+& {\scriptstyle{1\over2}} {\cal A} \ln \det \left[ I - \left\{ \left( D^\dagger - {\delta F\over \delta\phi}^\dagger \right)^{-1} g_2 \left( D - {\delta F\over \delta\phi} \right)^{-1} \left( (D \phi-F[\phi]) g_2^{-1} {\delta^2 F\over\delta\phi\;\delta\phi} \right) \right\} \right] \nonumber\\ &&- \left( \phi \to \phi_0 \right) + O( {\cal A} ^2). \end{eqnarray} This expression for the one-loop effective action is instructive. It is made up of two contributions whose origin and physics are quite different. On the one hand, the first term (the generalized Onsager-Machlup term) gives a contribution whose form is directly related to both the noise {\em shape factor} and the non-noisy part of the equation of motion, including non-linearities. On the other hand, the log-determinant term is proportional to the noise {\em amplitude} (which we have seen is the loop expansion parameter) and its specific form depends {\em also} on the structures of $D$ and $F[\phi]$, as well as on properties of the noise shape function. Therefore, noise plays a central role in the physics of the SPDE and, as will be discussed below, particularly in the nature of the ground state of the stochastic system described by equation (\ref{E:eom0}). \section{Effective Potential: One Loop} We now concentrate on field configurations that are homogeneous and static. For such field configurations the effective action reduces to a quantity known as the ``effective potential''. In this section we will {\em calculate} the effective potential, deferring the discussion of its physical {\em interpretation} (in the context of SPDEs) to the next section. The effective potential is defined as \begin{equation} {\cal V}[\phi;\phi_0] = {\Gamma[\phi;\phi_0] \over \Omega}, \end{equation} with $\phi$ a homogeneous and static field configuration, and $\Omega$ the volume of spacetime. The effective potential at one loop is given by \begin{eqnarray} {\cal V}[\phi;\phi_0] &=& {\scriptstyle{1\over2}} F^2[\phi] \left\{ \int {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; g_2^{-1} \right\} - {\scriptstyle{1\over2}} { {\cal A} \over\Omega} \ln \det \left( D - {\delta F\over \delta\phi} \right) - {\scriptstyle{1\over2}} { {\cal A} \over\Omega} \ln \det \left( D^\dagger - {\delta F\over \delta\phi}^\dagger \right) \nonumber\\ &&+ {\scriptstyle{1\over2}} { {\cal A} \over\Omega} \ln \det \left[ \left( D^\dagger - {\delta F\over \delta\phi}^\dagger \right) g_2^{-1} \left( D - {\delta F\over \delta\phi} \right) + F[\phi] \left\{ \int {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; g_2^{-1} \right\} {{\delta^2 F}\over{\delta\phi\;\delta\phi}} \right] \nonumber\\ &&- \left( \phi \to \phi_0 \right) + O( {\cal A} ^2). \label{E:ep0} \end{eqnarray} In order to turn this into a more tractable expression it is useful to introduce a frequency-momentum representation. First notice that \begin{eqnarray}\label{noisenorm} \int {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; g_2^{-1}(\vec x,t) &=& \int {{\mathrm{d}}^d \vec k \; {\mathrm{d}}\omega \over(2\pi)^{d+1}} \; {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; \tilde g_2^{-1}(\vec k,\omega) \; \exp[-i(\omega t - \vec k \cdot \vec x)] \nonumber\\ &=& \tilde g_2^{-1}(\vec k=\vec 0,\omega=0). \end{eqnarray} (It is clear from the formula for the one-loop effective potential equation (\ref{E:ep0}) that the above integral has to be finite, or rendered finite by appropriate renormalizations of the noise correlation function and the parameters it contains.) We next make use of the following identity valid for a translation invariant operator $X$ \begin{eqnarray} \ln \det X &=& \int {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; \int {\mathrm{d}}^d \vec k_1 \; {\mathrm{d}} \omega_1 \; \int {\mathrm{d}}^d \vec k_2 \; {\mathrm{d}} \omega_2 \; \langle \vec x,t | \vec k_1,\omega_1 \rangle \; \ln X(\vec k_1, \omega_1) \; \delta^d(\vec k_1,\vec k_2) \; \delta(\omega_1,\omega_2) \; \langle \vec k_2, \omega_2 |\vec x,t \rangle \nonumber\\ &=& \Omega \int {{\mathrm{d}}^d \vec k \; {\mathrm{d}} \omega\over (2\pi)^{d+1}} \ln X(\vec k, \omega). \end{eqnarray} Applying this to the one loop effective potential yields \begin{eqnarray} {\cal V}[\phi;\phi_0] &=& {\scriptstyle{1\over2}} F^2[\phi] \; \tilde g_2^{-1}(\vec k=\vec 0,\omega=0) - {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm{d}}^d \vec k \; {\mathrm{d}} \omega\over (2\pi)^{d+1}} \ln \left[ D(\vec k,\omega) - {\delta F\over \delta\phi} \right] - {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm{d}}^d \vec k \; {\mathrm{d}} \omega\over (2\pi)^{d+1}} \ln \left[ D^\dagger(\vec k,\omega) - {\delta F\over \delta\phi}^\dagger \right] \nonumber\\ &&+ {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm{d}}^d \vec k \; {\mathrm{d}} \omega\over (2\pi)^{d+1}} \ln \left[ \left( D^\dagger(\vec k,\omega) - {\delta F\over \delta\phi}^\dagger \right) \tilde g_2^{-1}(\vec k,\omega) \left( D(\vec k,\omega) - {\delta F\over \delta\phi} \right) + F[\phi] {\delta^2 F\over\delta\phi\;\delta\phi} \; \tilde g_2^{-1}(\vec k=\vec 0,\omega=0) \right] \nonumber\\ &&- \left( \phi \to \phi_0 \right) + O( {\cal A} ^2). \label{E:ep1} \end{eqnarray} (Note that $g_2$ plays two rather different roles above.) We now adopt the simplifying {\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=0,\omega=0). \end{equation} This is only a {\em convention}, not an additional restriction on the noise, since it only serves to give an absolute meaning to the normalization of the amplitude $ {\cal A} $. With these conventions, the one-loop effective potential can be written as \begin{eqnarray} {\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[ D(\vec k,\omega) - {\delta F\over \delta\phi} \right] - {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm{d}}^d \vec k \; {\mathrm{d}} \omega\over (2\pi)^{d+1}} \ln \left[ D^\dagger(\vec k,\omega) - {\delta F\over \delta\phi}^\dagger \right] \nonumber\\ &&+ {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm{d}}^d \vec k \; {\mathrm{d}} \omega\over (2\pi)^{d+1}} \ln \left[ \left( D^\dagger(\vec k,\omega) - {\delta F\over \delta\phi}^\dagger \right) \tilde g_2^{-1}(\vec k,\omega) \left( D(\vec k,\omega) - {\delta F\over \delta\phi} \right) + F[\phi] {\delta^2 F\over\delta\phi\;\delta\phi} \right] \nonumber\\ &&- \left( \phi \to \phi_0 \right) + O( {\cal A} ^2), \label{E:ep2} \end{eqnarray} which can be recast into \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^\dagger \over \delta\phi} \right) \left( D(\vec k,\omega) - {\delta F\over \delta\phi} \right)} \right] \nonumber\\ &&- \left( \phi \to \phi_0 \right) + O( {\cal A} ^2). \end{eqnarray} This formula is one of the central results of this paper. It shows that noise induced fluctuations modify the zero-loop piece of the potential in a way which is reminiscent of the situation in both statistical and quantum field theory. For example, in QFT one has~\cite{Weinberg-2,Rivers}: \begin{eqnarray} \label{E:general1} {\cal V}_{QFT}[\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} We see in equation (\ref{E:general}) that the ground state structure of the SPDE (which we will soon see is obtained by minimizing ${\cal V}[\phi;\phi_0]$) depends on both the noise correlations and the nonlinearities induced by the forcing term. We also see explicitly how the noise amplitude is essential in the competition between deterministic and stochastic effects. The major difference between the effective potential for SPDEs and QFT, lies in the fact that for SPDEs the scalar propagator of QFT is replaced with a propagator which has a more complex structure for the equivalent of the ``mass'' term. This difference owes to the causal structure of SPDEs. Notice also that for SPDEs one can naturally adapt the noise to be both the source of fluctuations {\em and} the regulator to keep the Feynman diagram expansion finite. This follows immediately by inspection of (\ref{E:general}) which shows that the (momentum and frequency dependent) noise shape function $\tilde g_2$ will affect the momentum and frequency behavior of the one-loop integral. The finiteness, divergence structure, and renormalizability of this integral will very much depend on the functional form of $\tilde g_2$. It is thus clear that we can use the noise shape function to regulate the integral, if we wish. \section{Interpretation} The {\em physical interpretation} of the effective action and the effective potential for SPDEs is considerably more subtle than that for the more usual QFTs. The situation is complicated by the fact that for a completely general SPDE it may not be meaningful to define a physical energy. Even when the SPDE is sufficiently special so that some physical notion of energy may be defined, the system may be subject to dissipation: The {\em physical} energy need not be conserved, even in the absence of noise. Thus the effective action and effective potential for SPDEs are not related to the physical energy. This means that {\em some} of the physical intuition built up from QFTs may be misleading and it becomes important to reassess the notion of effective action, and effective potential to see how much survives in the SPDE context. The great virtue of the effective action and effective potential in QFT is that they contain all the information regarding the ground state of the system and its fluctuations: From a knowledge of the effective potential one can ascertain under what conditions the system will display one degree of symmetry or another. It is essential that most of these properties carry over to the case of SPDEs, otherwise the effective action and effective potential would be mere mathematical constructs without physical relevance. Fortunately the key features do in fact carry over: (1) the stationary points of the effective action still correspond to stochastic expectation values of the fields in the absence of an external current; (2) the effective potential governs the probability that the {\em spacetime average} of the field takes on specific values; (3) even when the notion of physical energy is lacking we will see that there is a notion of quasi-energy for SPDEs, with the quasi-energy being a measure of the extent to which the system has been driven away from its non-stochastic (zero-noise) configuration; and (4) the one-loop effective action will be demonstrated to describe the (approximate) probability for an initial field configuration to evolve into some final (in the asymptotic sense) field configuration under the influence of the stochastic noise. \subsection{Equations of motion in the presence of fluctuations} If one makes use of the definition of the effective action as a Legendre transform, it is easy to see that \begin{equation} {\delta \Gamma[\phi;\phi_0] \over \delta \phi} = J[\phi], \end{equation} where $J[\phi]$ is that external current required in order that \begin{equation} \langle \phi[J] \rangle = \phi. \end{equation} In particular, by taking $J=0$ \begin{equation} {\delta \Gamma[\phi;\phi_0] \over \delta \phi} = 0 \qquad \Leftrightarrow \qquad \phi = \langle \phi[J=0] \rangle. \end{equation} Stationary points of the effective action occur at those (mean) field configurations which are zero-external-current stochastic expectation values of the fluctuating field. (Proof of this may be found for instance on page 65 of Weinberg~\cite{Weinberg-2}). It is important to realize that one never needs to invoke the notion of energy to obtain this result. The QFT interpretation of this result, which we now see extends to SPDEs, is that the effective action gives the equations of motion once fluctuations (noise) are taken into account. (This is a non-perturbative result, not limited to the one-loop approximation.) \subsection{Probability distribution for the spacetime average field} We have previously seen that the probability distribution for the fluctuating field, considered as a function over spacetime, to take on the value $\phi(\vec x,t)$ is given by the functional \begin{equation} {\bf P} [\phi] = {\cal P} [D\phi-F[\phi]] \; \sqrt{ {\cal J} {\cal J} ^\dagger}. \end{equation} Now suppose we coarse-grain, by looking at the spacetime average of the field $\phi$ as defined by \begin{equation} {\int_{\Omega_s\times T} \phi(\vec x,t) \; d^d\vec x \; dt \over \Omega_s T}, \end{equation} and ask: what is the probability that this spacetime average take on a specific numerical value $\bar\phi$? (For definiteness we impose periodic boundary conditions in space $\Omega_s$ and time $T$ and interpret $\Omega_s T$ as the volume of the spacetime box. This has the technical advantage that the partition function $Z[J]$ is then needed only for sources $J$ that are strictly independent of space and time.) The probability we are interested in is easily calculated to be \begin{eqnarray} \hbox{Prob}\left( \int_{\Omega_s\times T} {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; \phi(\vec x,t) = \bar\phi \; \Omega_s T\right) &=& \int ({\cal D} \phi) \; {\bf P} [\phi] \; \delta\left( \int_{\Omega_s \times T} {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; \phi(\vec x,t) - \bar\phi \; \Omega_s T \right) \\ &=& \int ({\cal D} \phi) \int d\lambda \; {\bf P} [\phi] \; \exp\left(i\lambda\left[ \int_{\Omega_s\times T} {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; \phi(\vec x,t) - \bar\phi \; \Omega_s T \right]\right) \\ &=& \int d\lambda \; Z[J(x)=i\lambda] \; \exp(-i\lambda\Omega_s T\;\bar\phi). \end{eqnarray} We now take the limit as $\Omega_s T$ becomes very large, and apply the method of stationary phase. By definition we have \begin{eqnarray} Z[J(x)=i\lambda] = \exp\left[ \Omega_s T \{ i\lambda \phi(\lambda) - {\cal V}[\phi(\lambda);\phi_0] /{\cal A} \} \right] \end{eqnarray} with the subsidiary condition \begin{equation} \left. \frac{\delta {\cal V}[\phi;\phi_0]}{\delta \phi} \right|_{\phi(\lambda)} = i \lambda {\cal A}. \end{equation} It is easy to demonstrate that \begin{eqnarray} \hbox{Prob}\left( \int_{\Omega_s\times T} {\mathrm{d}}^d \vec x \; {\mathrm{d}} t \; \phi(\vec x,t) = \bar\phi \; \Omega_s T \right) &\propto& \exp\left( -\Omega_s T\left[ {{\cal V}[\bar\phi;\phi_0]\over{\cal A}} + O\left({1\over\Omega_s T}\right) \right] \right). \end{eqnarray} Thus the effective potential governs the probability distribution of the spacetime average of the fluctuating field. Minima of the effective potential correspond to maxima of the probability density of the spacetime average field. The way we have set up the argument applies equally well to QFTs and SPDEs and makes no reference to the notion of physical energy. (This result is non-perturbative but approximate---it is not limited to one loop. If we take either the infinite volume or infinite time limits, then with probability one, the spacetime average field must equal one of the minima of the effective potential.) \subsection{Action and quasi-energy for SPDEs} Even though the physical energy may not be defined for arbitrary SPDEs, we nevertheless can demonstrate that there always exists a positive-semi-definite functional of field configurations, the tree-level action, and a related ``quasi-energy'', whose minima correspond to maxima of the probability distribution of field configurations. {From} the way the functional formalism has been set up, we can always define and calculate an effective action and an effective potential even if the underlying non-stochastic version of the partial differential equation does not arise from a Lagrangian formulation. We have already seen that the effective action has a natural interpretation in terms of the equations of motion once fluctuations are taken into account, and that the effective potential governs fluctuations in the spacetime average of the field. We now go one step further: We distinguish two concepts of ``energy'', the ``true physical energy'' and the ``quasi-energy'', and show that even if the physical energy is undefined (or possibly not useful due to dissipative effects), the quasi-energy is still a useful measure of the extent to which fluctuations modify the non-stochastic equations of motion. We start from our general SPDE (\ref{E:eom0}) \begin{equation} \label{E:noisy} D \phi = F[\phi] + \eta, \end{equation} and its non-stochastic version \begin{equation} \label{E:zero-noise} D\phi = F[\phi]. \end{equation} Sometimes this non-stochastic partial differential equation will arise from some Lagrangian, often it will not. Even if the PDE $(D\phi = F[\phi])$ does not arise from a Lagrangian, the results of this paper demonstrate that it is always possible to assign a tree-level action to the stochastic system: \begin{equation} {\cal S}_{ classical} = {\scriptstyle{1\over2}} \int \int dx\,dy\, (D \phi - F[\phi]) g_2^{-1} (D \phi - F[\phi]) \geq 0. \end{equation} This classical action is positive semidefinite, and has minima (which are equal to zero) at field configurations that satisfy the zero-noise equations of motion. This is most obvious for white noise, when the action is a perfect square, but the result is general. The noise two-point correlation function, [being an (infinite dimensional) covariance matrix], is by definition positive definite. Therefore its inverse is also positive definite and similarly the (infinite dimensional) matrix $g_2^{-1}$ is a positive definite operator. Thus this classical action (the generalized Onsager--Machlup action~\cite{Onsager-Machlup}) is always greater than or equal to zero. The classical action thus measures the extent to which a given field configuration fails to satisfy the zero-noise equations of motion; the measure of the deviation being weighted by the {\em shape} of the noise correlations. (In fact, if the amplitude ${\cal A}$ of the noise is set to zero, the action is identically equal to zero.) We now define the quasi-energy by \begin{equation} \label{quasiHam} {\cal S}[\phi] = \int E_{ quasi}[\phi] \; dt. \end{equation} We justify calling this object the quasi-energy by the fact that if we treat it as a Hamiltonian functional, and put the resulting object into the partition function of an equilibrium statistical field theory, we get the generating functional for all the correlation functions (ignoring ghost Jacobians for the moment). Explicitly, we can write \begin{equation} E_{ quasi}[\phi] = {\scriptstyle{1\over2}} \int \int d^d \vec x \; d^d \vec y \; dt' \; (D \phi - F[\phi]) g_2^{-1} (D \phi - F[\phi]). \end{equation} Note that the quasi-energy depends both on the PDE and on the shape of the noise correlation function. If the amplitude ${\cal A}$ of the noise is set to zero, this quasi-energy is conserved and is exactly equal to zero. This quasi-energy can be thought of as a nonlinear generalization of the Onsager--Machlup ``energy'' to arbitrary Gaussian noise\footnote{The particular label one chooses to apply to this quantity is not important as long as one bears carefully in mind that this ``energy'' need not be the physical energy.}. If we now restrict ourselves to homogeneous and static fields, and consider the effective potential as defined above, then by the procedures used in quantum and stochastic field theories, the effective potential (multiplied by the volume of space) is the stochastic expectation value of the quasi-energy $\langle E_{ quasi}[\phi]\rangle$ in the {\em presence} of the noise induced fluctuations, and subject to the constraint $\langle\phi\rangle = \phi$. A proof of this result is provided on pages 72--73 of Weinberg~\cite{Weinberg-2}. Though that proof is phrased in a Lorentzian-signature QFT language, it readily carries over to Euclidean-signature equilibrium statistical field theory. Once the physical energy is replaced by the quasi-energy, the proof can be extended to SPDEs as well. We have been able to show that minima of the effective potential also minimize the quasi-energy, and therefore the noise-induced deviations from the zero-noise equations of motion. \subsection{Transition probabilities} What is the probability that a certain initial field configuration $\phi_i(\vec x)$ at time $t_i$ evolves into a final field configuration $\phi_f(\vec x)$ at time $t_f$? We have already developed the appropriate machinery to address this question. Indeed \begin{eqnarray} \hbox{Prob}(\phi_f(\vec x),t_f; \phi_i(\vec x), t_i) &\propto& \int ({\cal D}\eta) \; {\cal P} [\eta] \; \delta[\phi_{ soln}(\vec x, t_i;\eta) - \phi_i(\vec x)]\; \delta[\phi_{ soln}(\vec x, t_f;\eta) - \phi_f(\vec x)] \nonumber\\ &\propto& \int ({\cal D}\eta) \; ({\cal D} \phi) \; {\cal P} [\eta] \; \delta[\phi_{ soln}(\vec x, t;\eta) - \phi(\vec x,t)]\; \delta[\phi(\vec x, t_i) - \phi_i(\vec x)]\; \delta[\phi(\vec x, t_f) - \phi_f(\vec x)] \nonumber\\ &\propto& \int ({\cal D}\eta) \; ({\cal D} \phi) \; {\cal P} [\eta] \; \delta[ D\phi - F[\phi] - \eta] \; \sqrt{ {\cal J J}^\dagger } \; \delta[\phi(\vec x, t_i) - \phi_i(\vec x)]\; \delta[\phi(\vec x, t_f) - \phi_f(\vec x)] \nonumber\\ &\propto& \int ({\cal D} \phi) \; {\cal P} [D\phi - F[\phi]] \; \sqrt{ {\cal J J}^\dagger } \; \delta[\phi(\vec x, t_i) - \phi_i(\vec x)]\; \delta[\phi(\vec x, t_f) - \phi_f(\vec x)] \nonumber\\ &\propto& \int ({\cal D} \phi) \; {\bf P} [\phi] \; \delta[\phi(\vec x, t_i) - \phi_i(\vec x)]\; \delta[\phi(\vec x, t_f) - \phi_f(\vec x)] \nonumber\\ &\propto& \int ({\cal D} \phi) \; \exp\left(-{\cal S}[\phi]/{\cal A}\right) \; \sqrt{ {\cal J J}^\dagger } \; \delta[\phi(\vec x, t_i) - \phi_i(\vec x)]\; \delta[\phi(\vec x, t_f) - \phi_f(\vec x)] \nonumber\\ &\propto& \int_{\phi(\vec x, t_i)=\phi_i(\vec x)}^{\phi(\vec x, t_f)=\phi_f(\vec x)} \; ({\cal D} \phi) \; \exp\left(-{\cal S}[\phi]/{\cal A}\right) \; \sqrt{ {\cal J J}^\dagger }. \end{eqnarray} This is formally identical to the formula usually encountered in equilibrium statistical field theory, and everything so far is non-perturbatively correct. Now take a saddle point approximation: Find an interpolating field $\phi_{ int}(\vec x,t)$ that minimizes ${\cal S}[\phi]$ and interpolates from $\phi_i(\vec x)$ to $\phi_f(\vec x)$. Perform the Gaussian integral about the saddle point. Then by definition of the one-loop effective action \begin{equation} \hbox{Prob}(\phi_f(\vec x),t_f; \phi_i(\vec x), t_i) \approx \exp\left[-{\Gamma[\phi_{ int}]\over{\cal A}}\right]. \end{equation} This is only a one-loop result, but it demonstrates that the effective action for SPDEs inherits many of the important features of the effective action for QFTs. \subsection{Summary} {From} the above, we see that the effective action and effective potential for SPDEs exhibit many of the key features of the effective action and effective potential of QFTs. This is important because it guarantees that not only is it relatively easy to calculate the one-loop effective potential, but also it is useful to do so: As is the case for QFTs, minima of the effective potential for SPDEs provide information about expectation values of the fields. The effective action also provides information about fluctuations in spacetime averaged fields, it gives information about the noise-induced deviations from the non-stochastic equations of motion, and it governs the transition probabilities whereby initial field configurations evolve to final field configurations. Thus, both the effective potential and the effective action are as useful for SPDEs as it is for QFTs. Furthermore, as demonstrated in recent work by Alexander and Eyink~\cite{Eyink1,Eyink2,Eyink3}, the effective potential is also a useful tool in a strong noise regime far from equilibrium. The major difference between those papers and our own formalism is that they work within the MSR approach. They also focus on strong noise regimes, while we emphasize that for many purposes a one-loop calculation is both computationally efficient and quite sufficient to extract many key features of the physics of the system. The two approaches are complementary, and where they overlap, they are in complete agreement. \section{Discussion} In this paper we have developed a general and powerful formalism applicable to arbitrary SPDEs. We have shown how to convert {\em arbitrary} correlation functions associated with {\em arbitrary} SPDEs into functional integrals. (And for this first step the noise does not have to be Gaussian.) For Gaussian noise (not necessarily translation invariant) we have carried the formalism further, setting up the basic ingredients needed for Feynman diagram expansions with the noise amplitude serving as the loop-counting parameter, and defining a non-perturbative effective action in analogy with QFT. We hope to have convinced the reader that the ``direct approach'' developed in this paper is both useful and complementary to the more traditional MSR formalism \cite{MSR,De-Dominicis-Peliti,Zinn-Justin}. Some questions can more profitably be asked and answered in this ``direct'' formalism. For instance, the fact that the noise amplitude is the loop-counting parameter is easy to establish in this ``direct'' formalism, but appears to have no analog result in the MSR formalism. The effective action gives rise naturally to the concept of an effective potential, a powerful construct well known and studied within the QFT context, where it serves to classify and compute ground states and allows one to investigate symmetry properties and patterns of symmetry breaking (both spontaneous and dynamic). An analogous construct can also be defined and calculated for stochastic field theories based on SPDEs, and we have done so in this work. However, for arbitrary SPDEs, such as those contemplated here, the notion of ground state and effective potential must be approached with extra care and their physical interpretation clarified. We have taken pains to do so, establishing that the minimum of the construct we call the effective potential corresponds to solving the full equations of motion (for homogeneous and static field configurations) in the presence of noise. We feel that the most interesting result of this analysis is a general formula for the one-loop effective potential for {\em any} SPDEs subject to translation-invariant Gaussian noise. This is still an extremely broad class of problems, and in a pair of companion papers we will specialize this analysis to two particular cases. First, we discuss the noisy Burgers equation (KPZ equation), where the effective potential approach immediately leads us to such interesting observations as the existence of dynamical symmetry breaking (DSB) and the Coleman--Weinberg mechanism~\cite{HMPV-kpz}. Second, we discuss the reaction--diffusion-decay system, and explicitly calculate the renormalized effective potential for 1, 2, and 3 spatial dimensions~\cite{HMPV-rdd}. These are issues that are extremely difficult to address using the MSR approach. \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). This research is supported in part by the Department of Energy under contract W-7405-ENG-36 (CMP). Additionally, MV wishes to acknowledge support from the Spanish Ministry of Education and Culture through the Sabbatical Programme, 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.
\subsection*{Introduction} This paper begins with the problem of determining the self-similar images of certain lattices and $\mathbb Z$-modules in four dimensions and ends in the enchanting garden of Coxeter groups, the arithmetic of several quaternion rings, and the asymptotics of their associated zeta functions. The main results appear in Theorems \ref{thm-lattices} and \ref{thm-modules}. The symmetries of crystals are of fundamental physical importance and, along with the symmetries of lattices, have been studied by mathematicians, crystallographers and physicists for ages. The recent interest in {\em quasicrystals}, which are non-crystallographic yet still highly ordered structures, has naturally led to speculation about the role of symmetry in this new context. Here, however, it is apparent that a different set of symmetry concepts is appropriate, notably because translational symmetry is either entirely lacking or at least considerably restricted in scope. One of the most obvious features of quasicrystals is their tendency to have copious inflationary self-similarity. Thus instead of groups of isometries, one is led to semi-groups of self-similarities that map a given (infinite) point set into an inflated copy lying within itself. Ordinary point symmetries then show up as a small part of this, namely as the ``units'', i.e. as the (maximal subgroups of) invertible elements. The importance of self-similarities is well-known, and has also been used to gain insight into the colour symmetries of crystals \cite{Schwarz1,Schwarz2} and, more recently, of quasicrystals \cite{Baake97,BM1,Ron}. The focus in the latter cases was on highly symmetric examples in the plane and in 3-space since they are obviously of greatest physical importance. In this contribution, we extend the investigation of self-similarities to certain exceptional examples in 4-space, namely the hypercubic lattices (of which there are two, represented by the primitive hypercubic lattice $\mathbb Z^4$ and the root lattice $D^{}_4$ or, equivalently, its weight lattice $D^*_4$) and the icosian ring, seen as the $\mathbb Z$-span of the root system of the non-crystallographic Coxeter group $H_4$, see \cite{CS,Humphreys,Cox80} for notation and background material. The case of lattices has recently also been investigated by Conway, Rains and Sloane \cite{CRS} who are generally interested in the question under which conditions similarity sublattices of a given index exist. Their methods are complementary to ours, and more general, but do not seem to give direct access to the full combinatorial problem which we can solve here. It is useful to digress briefly to discuss the role of $H_4$ in the context of aperiodic order. It is a remarkable fact that the non-crystallographic point symmetries relevant to essentially all known physical quasicrystals are actually Coxeter groups, namely the dihedral groups $I_2(k)$, $k = 5,8,10,12$ and the icosahedral group $H_3$ \cite{BJKS}. Apart from the remaining infinite series of dihedral groups, the only indecomposable non-crystallographic Coxeter group is $H_4$ of order 14400, which is the most interesting of them all. In spite of being four-dimensional in nature, there are several good physical reasons to explore the symmetry expressed by this group, and because of its connections with exceptional objects in mathematics, including the root lattice $E_8$, there are good mathematical reasons, too. {}From the point of view of quasicrystals, $H_4$ may be viewed as the top member of the series $H_2 := I_2(5) \subset H_3 \subset H_4$ of which the first two have been the subject of great attention, see \cite{Kramer} and references therein, while $H_4$ appears as symmetry group of the Elser-Sloane quasicrystal \cite{ES}. Mathematically, this family belongs together and $H_4$ is the natural parent of the others. Now, $H_4$ has a natural quaternionic interpretation which arises as follows. The group of norm $1$ units of the real quaternion algebra is easily identifiable with $\mbox{SU}(2)$, see \cite{Koecher} for background material and notation. Using the $2$-fold cover of $\mbox{SO}(3)$ by $\mbox{SU}(2)$ we can find (in many ways) a $2$-fold cover of the icosahedral group inside $\mbox{SU}(2)$. This is the binary icosahedral group $I$ of order $120$ \cite[p.\ 69]{CM}. The point set $I$ is a beautiful object, namely the set of vertices of the exceptional regular $4$-polytope called the $600$-cell\footnote{The 600-cell and its dual, the 120-cell, are two of the three exceptional regular polytopes in 4-space \cite[p.\ 292]{Cox73}. The remaining one, the 24-cell, also occurs, later in this paper. Beyond 4 dimensions, the only regular polytopes are the simplices, the hypercubes, and their duals, the cross-polytopes (or hyperoctahedra).} \cite[Ch.\ 22]{Cox80} (under the standard topology of $\mathbb R^4$ carried by the quaternion algebra). The $\mathbb Z$-span of $I$ is a ring $\mathbb I$, dubbed by Conway the ring of {\em icosians}. This ring is closed under complex conjugation and under left and right multiplications by elements of $I$. The group of symmetries of $\mathbb I$ so obtained is isomorphic to $H_4$ acting as a reflection group in $\mathbb R^4$. The ring $\mathbb I$ is itself quite a remarkable object. It is naturally a rank $4$-module over $\mathbb Z[\tau]$, $\tau := (1 + \sqrt 5)/2$, and a rank $8$-module over $\mathbb Z$ (so it is certainly dense in the ambient space $\mathbb R^4$). In fact, as an aside, it has a canonical interpretation as the root lattice of type $E_8$ with $I \cup \tau I$ making up the $240$ roots of $E_8$. Restricting to the pure quaternions puts us in the $3$-dimensional icosahedral case, and by further restriction we can get the $H_2$ situation. Now we can state our problem for the icosian case. We have pointed out that the additve group $\mathbb I$ has a large finite group of rotational symmetries coming from the left and right multiplications by elements of $I$ -- namely $120^2/2=7200$ such symmetries. But we have seen that in the study of quasicrystals we have to pay attention to self-similarities, too. So we are now also interested in rotation-inflations of $\mathbb R^4$ that map $\mathbb I$ into, but not necessarily onto, itself. Each such self-similarity maps $\mathbb I$ onto some submodule of finite index, and our question is to determine these images and to count the number of different similarity submodules of a given index. This leads us to introduce a suitable Dirichlet series generating function which encodes the counting information and its asymptotic properties all at once, and indeed determining its exact form is a number theoretical problem that depends heavily on the fact that $\mathbb I$ can be interpreted as a maximal order in the split quaternionic algebra over the quadratic field $\mathbb Q(\mbox{\small $\sqrt{5}$})$. The other situation that we wish to discuss in this paper is crystallographic in origin, but it nevertheless is very much the same problem. It is well known that there are two hypercubic lattices in $4$-space \cite{Schwarz1,Brown}, namely the primitive and the centred one (face-centred and body-centred are equivalent in $4$-space by a similarity tranformation). Let us take $\mathbb Z^4$ and the root lattice $D_4$ as suitable representatives. Note that they have different holohedries, namely one of order 1152 (denoted by 33/16 in \cite[Fig.\ 7]{Brown}) for $D_4$, which coincides with the automorphism group of the underlying root system, and an index 3 subgroup (denoted by 32/21 in \cite[Fig.\ 7]{Brown}) for $\mathbb Z^4$. Due to the previous remark, we may take the weight lattice $D^*_4$ instead of $D_4$ if we wish, and we will frequently do so. Given any of these cases, we want to determine the Dirichlet series generating function for the sublattices that are self-similar images of it. What makes this situation tractable is that, parallel to the icosian case, there is a highly structured algebraic and arithmetic object in the background, namely Hurwitz' ring of integral quaternions \cite{Hurwitz,Cox46}. In our setting, it is $\mathbb J=D^*_4$, and it is again a maximal order, this time of the quaternionic algebra over the rationals, $\mathbb Q$. The results for the Hurwitzian and icosian cases are striking in their similarity. Another example is that of the maximal order in $\mathbb H(\mathbb Q(\mbox{\small $\sqrt{2}$}))$ the treatment of which equals that of $\mathbb I$ whence we only state the results. The structure of the paper is as follows. In the next Section, we set the scene by collecting some methods and results from algebra, and algebraic number theory in particular. This will be done in slightly greater detail than necessary for a mathematical audience, but since there is also considerable interest in this type of problem from the physics community, we wish to make the article more self-contained and readable this way. The two following Sections give the results, first for lattices, and then for modules. We close with a brief discussion of related aspects and provide an Appendix with material on the asymptotics of arithmetic functions defined through Dirichlet series. \subsection*{Preliminaries and Recollections} We shall need a number of results from algebraic number theory, both commutative and non-commutative. {}First of all, we need, of course, the arithmetic of $\mathbb Z$, the ring of integers in the field $\mathbb Q$. All ideals of $\mathbb Z$ are principal, and they are of the form $\mathfrak a = m\mathbb Z$ with $m\in\mathbb Z$. If $\mathfrak a\neq 0$, the index is $[\mathbb Z:\mathfrak a]=|m|$. The corresponding zeta function, which can be seen as the Dirichlet series generating function for the number of non-zero ideals of a given index, is Riemann's zeta function itself \cite{Apostol} \begin{equation} \label{Riemann} \zeta(s) \; = \; \sum_{\mathfrak a\subset\mathbb Z}\frac{1}{[\mathbb Z:\mathfrak a]^s} \; = \; \sum_{m=1}^{\infty} \frac{1}{m^s} \; = \; \prod_{p\in{\cal P}} \frac{1}{1-p^{-s}} \, . \end{equation} Here, ${\cal P}$ denotes the set of (rational) primes, and the second representation of the Riemann zeta function is its {\em Euler product expansion}. It is possible because the number of ideals of index $m$ is a multiplicative arithmetic function, a situation that we shall encounter throughout the article. Next, we need the analogous objects for the real quadratic field $\mathbb Q(\mbox{\small $\sqrt{5}$})=\mathbb Q(\tau)$. The ring of integers turns out to be \begin{equation} \mathbb Z[\tau] \; = \; \{ m + n \tau \mid m,n \in \mathbb Z \} \end{equation} where $\tau = (1+\mbox{\small $\sqrt{5}$})/2$ is the fundamental unit of $\mathbb Z[\tau]$, i.e.\ all units are obtained as $\pm\tau^m$ with $m\in\mathbb Z$. Again, $\mathbb Z[\tau]$ is a principal ideal domain, and hence a unique factorization domain \cite[ch.\ 15.4]{Hardy}. The zeta function is the Dedekind zeta function \cite[\S 11]{Zagier} defined by \begin{equation} \zeta_{\mathbb Q(\tau)}^{}(s) \; = \; \sum_{\mathfrak a\subset\mathbb Z[\tau]} \frac{1}{[\mathbb Z[\tau]:\mathfrak a]^s} \; = \; \sum_{m=1}^{\infty} \frac{a(m)}{m^s} \end{equation} where $\mathfrak a$ runs through the non-zero ideals of $\mathbb Z[\tau]$ and $[\mathbb Z[\tau]:\mathfrak a]$ is the norm of $\mathfrak a$. If $\mathfrak a = \alpha\mathbb Z[\tau]$, it is given by \begin{equation} [\mathbb Z[\tau]:\mathfrak a] \; = \; |\mbox{N}(\alpha)| \; = \; |\alpha \alpha'| \end{equation} where $'$ denotes algebraic conjugation in $\mathbb Q(\tau)$, defined by $\tau \mapsto 1-\tau$. Explicitly, the zeta function reads (see the Appendix for details): \begin{eqnarray} \label{zeta2} \zeta_{\mathbb Q(\tau)}^{}(s) &\!\!\! = &\!\!\! \frac{1}{1-5^{-s}} \cdot \prod_{p \equiv \pm 1 \; (5)} \frac{1}{(1-p^{-s})^2} \cdot \prod_{p \equiv \pm 2 \; (5)} \frac{1}{1-p^{-2s}} \\ &\!\!\! = &\!\!\! \mbox{\small $ 1+\frac{1}{4^s}+\frac{1}{5^s}+\frac{1}{9^s}+\frac{2}{11^s}+ \frac{1}{16^s}+\frac{2}{19^s}+\frac{1}{20^s}+\frac{1}{25^s}+ \frac{2}{29^s}+\frac{2}{31^s}+ \frac{1}{36^s}+\frac{2}{41^s} + \cdots $} \nonumber \end{eqnarray} As before, $a(m)$ is a multiplicative arithmetic function, i.e.\ $a(mn)=a(m)a(n)$ for coprime $m,n$. It is thus completely specified by its value for $m$ being a prime power, and from the Euler product in (\ref{zeta2}) one quickly derives that $a(5^r)=1$ (for $r\geq 0$). Then, for primes $p\equiv\pm 2$ $(5)$, one obtains $a(p^{2r+1})=0$ and $a(p^{2r})=1$, while for primes $p\equiv\pm 1$ $(5)$, the result is $a(p^r) = r+1$. One benefit of relating the numbers $a(m)$ to zeta functions with well-defined analytic behaviour is that one can rather easily determine the asymptotic behaviour of $a(m)$ from the poles of the zeta function, see the Appendix for a summary. In this case, the function $a(m)$ is constant on average, the constant being the residue of $\zeta^{}_{\mathbb Q(\tau)}(s)$ at its right-most pole, $s=1$. Explicitly, we get \begin{equation} \label{asymp1} \lim_{N\to\infty} \frac{1}{N} \sum_{m=1}^{N} a(m) \; = \; \mbox{res}_{s=1} \, \zeta^{}_{\mathbb Q(\tau)}(s) \; = \; \frac{2\log(\tau)}{\sqrt{5}} \; \simeq \; 0.430409 \, . \end{equation} We shall also need the zeta function of the quadratic field $\mathbb Q(\mbox{\small $\sqrt{2}$})$, where $\mathbb Z[\mbox{\small $\sqrt{2}$}]$ is the corresponding ring of integers, and $1+\mbox{\small $\sqrt{2}$}$ its fundamantal unit \cite{Hardy}. The zeta function reads \begin{eqnarray} \label{zeta3} \zeta_{\mathbb Q({\scriptscriptstyle \sqrt{2}})}^{}(s) &\!\!\! = &\!\!\! \frac{1}{1-2^{-s}} \cdot \prod_{p \equiv \pm 1 \; (8)} \frac{1}{(1-p^{-s})^2} \cdot \prod_{p \equiv \pm 3 \; (8)} \frac{1}{1-p^{-2s}} \\ &\!\!\! = &\!\!\! \mbox{\small $ 1+\frac{1}{2^s}+\frac{1}{4^s}+\frac{2}{7^s}+\frac{1}{8^s}+ \frac{1}{9^s}+\frac{2}{14^s}+\frac{1}{16^s}+\frac{2}{17^s}+ \frac{1}{18^s}+\frac{2}{23^s}+ \frac{1}{25^s}+\frac{2}{28^s} + \cdots $} \nonumber \end{eqnarray} The coefficients are $a(2^r)=1$, $a(p^r)=r+1$ for $p\equiv \pm 1$ (8) and $a(p^r)=0$ resp.\ $1$ for $p\equiv \pm 3$ (8) and $r$ odd resp.\ even. The asymptotic behaviour is given by $\lim^{}_{N\to\infty} {1\over N}\sum_{m\leq N} a(m) = \log(1+\mbox{\small $\sqrt{2}$})/\mbox{\small $\sqrt{2}$} \simeq 0.623225$. Let us now move to the non-commutative results we shall need. We will be concerned with the quaternionic algebra $\mathbb H({K})$, mainly over the field ${K}=\mathbb Q$ or over ${K}=\mathbb Q(\tau)$. The case of ${K}=\mathbb Q(\mbox{\small $\sqrt{2}$})$ is treated more as an aside. In all cases, we are interested in the corresponding ring of integers, these being the {\em Hurwitzian ring} $\,\mathbb J$, the {\em icosian ring} $\,\mathbb I$, and the {\em cubian ring} $\,\mathbb K$. These are maximal orders in their respective quaternionic algebras \cite{Hurwitz,Vigneras}. Let us first consider $\mathbb J=D^*_4$. In terms of the standard basis\footnote{The defining relations \cite{Cox46} are: $\bm{i}^2=\bm{j}^2=\bm{k}^2=\bm{ijk}=-1$.} $1,\bm{i},\bm{j},\bm{k}$ of $\mathbb H(\mathbb Q)$, $\mathbb J$ consists of the points $(x_0,x_1,x_2,x_3)$ whose coordinates $x_i$ either all lie in $\mathbb Z$ or all in $\mathbb Z+\frac{1}{2}$. Though non-commutative, $\mathbb J$ is still a principal ideal domain, i.e.\ all left-ideals (and also all right-ideals) are principal \cite{Hurwitz}. Consequently, we have unique factorization up to units, the units being the 24 elements obtained from $(\pm 1,0,0,0)$ plus permutations and from ${1\over 2}(\pm 1,\pm 1,\pm 1,\pm 1)$. These 24 units form a group, $\mathbb J^{\times}$, that is isomorphic to the binary tetrahedral group \cite[p.\ 69]{CM}. The number of non-zero left-ideals (or, equivalently, of non-zero right-ideals) can now be counted explicitly, which results in the corresponding zeta function $ \zeta^{}_{\mathbb J}(s) \; = \; \sum_{\mathfrak a} \frac{1}{[\mathbb J:\mathfrak a]^s}$, where $\mathfrak a$ runs over the non-zero left ideals of $\mathbb J$, see \cite[Sec.\ VII, {\S} 8 and {\S} 9]{Deuring} for details on zeta functions of quaternionic algebras. The result is \cite[{\S} 63, A.~15]{Scheja}: \begin{itemize} \item The zeta function of Hurwitz' ring of integer quaternions, $\mathbb J$, is given by\footnote{Note that the formula given in the first line after \cite[Eq.\ 3.17]{Baake} contains a misprint in the prefactor.} \begin{equation} \label{Hurwitz-zeta} \zeta^{}_{\mathbb J}(s) \; = \; (1-2^{1-2s}) \cdot \zeta(2s) \, \zeta(2s-1) \, . \end{equation} \end{itemize} Using (\ref{Riemann}), one can easily determine the first few terms \begin{equation} \zeta^{}_{\mathbb J}(s) \; = \; \mbox{\small $ 1+\frac{1}{4^s}+\frac{4}{9^s}+\frac{1}{16^s}+\frac{6}{25^s}+ \frac{4}{36^s}+\frac{8}{49^s}+\frac{1}{64^s}+\frac{13}{81^s}+ \frac{6}{100^s}+\frac{12}{121^s}+\frac{4}{144^s}+ \cdots $ } \end{equation} The possible indices of ideals are the squares of integers. It is thus convenient to write the Dirichlet series as \begin{equation} \label{coeff1} \zeta^{}_{\mathbb J}(s) \; = \; \sum_{m=1}^{\infty} \frac{a^{}_{\mathbb J}(m)}{m^{2s}} \end{equation} so that $a^{}_{\mathbb J}(m)$ is actually the number of left-ideals of index $m^2$ (rather than $m$). This then results in $a^{}_{\mathbb J}(2^r)=1$ (for $r\geq 0$) and in $a^{}_{\mathbb J}(p^r)=(p^{r+1}-1)/(p-1)$ for odd primes. Let us add that $a^{}_{\mathbb J}(m)$ is also the sum of the odd divisors of $m$, see \cite[sequence M 3197]{SP} or \cite[sequence A 000593]{Sloane}. Let us again briefly comment on the asymptotic behaviour. In this case, the average of $a^{}_{\mathbb J}(m)$ grows linearly with $m$, i.e.\ \begin{equation} \label{asymp2} \frac{1}{N} \sum_{m=1}^{N} a^{}_{\mathbb J}(m) \; \sim \; \frac{\pi^2}{24} N \quad \quad \quad (N \rightarrow\infty) \end{equation} where the coefficient is half the residue of $\zeta^{}_{\mathbb J}(s)$ at its right-most pole, $s=1$. This can easily be calculated from the details provided in the Appendix. Note that the linear growth of the average of $a^{}_{\mathbb J}(m)$ stems from the definition used in (\ref{coeff1}). With the usual definition (i.e.\ with denominators $m^s$ rather than $m^{2s}$), the average would tend to a constant, as in (\ref{asymp1}). Now, let us consider the analogous situation, with $\mathbb I$ being a maximal order in the algebra $\mathbb H(\mathbb Q(\tau))$. Again, all left-ideals (and all right-ideals) of $\mathbb I$ are principal, and we also get unique factorization up to units again \cite{Vigneras}. The unit group $\mathbb I^{\times}$ of $\mathbb I$ consists of the 120 elements of the binary icosahedral group $I$ inside $\mathbb I$. Taking the unit quaternions $(1,0,0,0)$, ${1\over 2}(1,1,1,1)$, ${1\over 2}(\tau,1,-1/\tau,0)$ together with all even permutations and arbitrary sign flips results in an explicit choice of the group $I$, and hence of $\mathbb I$. Again defining the zeta function for one-sided ideals of $\mathbb I$ we have \begin{itemize} \item The zeta function of the icosian ring, $\mathbb I$, is \begin{equation} \label{icosian-zeta} \zeta^{}_{\mathbb I}(s) \; = \; \zeta^{}_{\mathbb Q(\tau)}(2s) \, \zeta^{}_{\mathbb Q(\tau)}(2s-1) \, . \end{equation} \end{itemize} This result follows from \cite[Ch.\ III, Prop.\ 2.1]{Vigneras}. The first few terms of this series read \begin{equation} \label{ico-zeta-2} \zeta^{}_{\mathbb I}(s) \; = \; \mbox{\small $ 1+\frac{5}{16^s}+\frac{6}{25^s}+\frac{10}{81^s}+ \frac{24}{121^s}+\frac{21}{256^s}+\frac{40}{361^s}+ \frac{30}{400^s}+\frac{31}{625^s}+\frac{60}{841^s}+ \frac{64}{961^s}+ \cdots $ } \end{equation} The possible indices ($=$ denominators) are the squares of integers that are representable by the quadratic form $x^2 + x y - y^2$, i.e.\ of integers all of whose prime factors congruent to 2 or 3 (mod 5) occur with even exponent only. Using a definition analogous to (\ref{coeff1}) above, the coefficient $a^{}_{\mathbb I}(m)$ is again a multiplicative arithmetic function. It is given by $a^{}_{\mathbb I}(5^r) = (5^{r+1}-1)/4$ (for $r\geq 0$), and, for primes $p\equiv\pm 2$ $(5)$, by $a^{}_{\mathbb I}(p^{2r+1})=0$ and $a^{}_{\mathbb I}(p^{2r})=(p^{2r+2}-1)/(p^2-1)$. {}Finally, for $p\equiv\pm 1$ $(5)$, one finds $a^{}_{\mathbb I}(p^r) = \sum_{l=0}^{r} (l+1) (r-l+1) p^l$. It is now listed as \cite[sequence A 035282]{Sloane}. The asymptotic behaviour of $a^{}_{\mathbb I}(m)$ is similar to that of $a^{}_{\mathbb J}(m)$ above, and we obtain \begin{equation} \label{asymp3} \frac{1}{N} \sum_{m=1}^{N} a^{}_{\mathbb I}(m) \; \sim \; \frac{2\pi^4\log(\tau)}{375} N \; \simeq \; 0.249997 \cdot N \quad \quad \quad (N \rightarrow\infty) \end{equation} where the slope is again half the residue of $\zeta^{}_{\mathbb I}(s)$ at its right-most pole, $s=1$, see the Appendix for details. Very similar is the situation of the ring $\mathbb K$ in $\mathbb H(\mathbb Q(\mbox{\small $\sqrt{2}$}))$, generated as the $\mathbb Z[\mbox{\small $\sqrt{2}$}]$-span of the basis $\{1, (1+\bm{i})/\mbox{\small $\sqrt{2}$}, (1+\bm{j})/\mbox{\small $\sqrt{2}$}, (1+\bm{i}+\bm{j}+\bm{k})/2\}$. The unit group $\mathbb K^{\times}$ is the binary octahedral group of order 48 \cite[p.\ 69]{CM}, and the symmetry group of $\mathbb K$ contains that of $\mathbb J$ as an index 2 subgroup. $\mathbb K$ is again a maximal order and a principal ideal domain \cite{Vigneras}. The zeta function of $\mathbb K$ reads \cite[Ch.\ III, Prop.\ 2.1]{Vigneras} \begin{eqnarray} \label{cubian-zeta} \zeta^{}_{\mathbb K}(s) & \!\!\! = &\!\!\! \zeta^{}_{\mathbb Q({\scriptscriptstyle \sqrt{2}})}(2s) \, \zeta^{}_{\mathbb Q({\scriptscriptstyle \sqrt{2}})}(2s-1) \\ &\!\!\! = &\!\!\! \mbox{\small $ 1+\frac{3}{4^s}+\frac{7}{16^s}+\frac{16}{49^s}+\frac{15}{64^s}+ \frac{10}{81^s}+\frac{48}{196^s}+\frac{31}{256^s}+ \frac{36}{289^s}+\frac{30}{324^s}+\frac{48}{529^s}+ \frac{26}{625^s} + \cdots $} \nonumber \end{eqnarray} and further details can be worked out in complete analogy to the icosian case. Let us now briefly describe how the quaternions enter our (mainly geometric) picture, and how they provide a parametrization of $\mbox{(S)O}(4) = \mbox{(S)O}(4,\mathbb R)$, see \cite{Koecher} for details. The key is that pairs of quaternions in $\mathbb H(\mathbb R)$, i.e.\ quaternions $\bm{q}=(q^{}_0,q^{}_1,q^{}_2,q^{}_3)$ as written in the standard basis $1,\bm{i},\bm{j},\bm{k}$ of the quaternion algebra over $\mathbb R$, induce an action on vectors of $\mathbb R^4$ via \begin{equation} \label{action} M(\bm{q}^{}_1,\bm{q}^{}_2) \bm{x}^t \; = \; \bm{q}^{}_1 \bm{x} \, \overline{\bm{q}}^{}_2 \end{equation} where $M(\bm{q}^{}_1,\bm{q}^{}_2)\in \mbox{Mat}(4,\mathbb R)$ and $\bm{x}^t$ is $\bm{x}$ written as a column vector in $\mathbb R^4$. Evidently, for nonzero quaternions $\bm{q}^{}_1,\bm{q}^{}_2,\bm{r}^{}_1,\bm{r}^{}_2$ we have $M(\bm{q}^{}_1,\bm{q}^{}_2) = M(\bm{r}^{}_1,\bm{r}^{}_2)$ if and only if $\bm{r}^{}_1 = a \bm{q}^{}_1$ and $\bm{r}^{}_2 = a^{-1}\bm{q}^{}_2$ for some $a \in \mathbb R \backslash \{0\}$. With $\bm{q}^{}_1 = (a,b,c,d)$ and $\bm{q}^{}_2 = (t,u,v,w)$, the matrix $M=M(\bm{q}^{}_1,\bm{q}^{}_2)$ reads explicitly \begin{equation} \label{matrix} \left( \begin{array}{rrrr} at\!+\!bu\!+\!cv\!+\!dw & \!-\!bt\!+\!au\!+\!dv\!-\!cw & \!-\!ct\!-\!du\!+\!av\!+\!bw & \!-\!dt\!+\!cu\!-\!bv\!+\!aw \\ bt\!-\!au\!+\!dv\!-\!cw & at\!+\!bu\!-\!cv\!-\!dw & \!-\!dt\!+\!cu\!+\!bv\!-\!aw & ct\!+\!du\!+\!av\!+\!bw \\ ct\!-\!du\!-\!av\!+\!bw & dt\!+\!cu\!+\!bv\!+\!aw & at\!-\!bu\!+\!cv\!-\!dw & \!-\!bt\!-\!au\!+\!dv\!+\!cw \\ dt\!+\!cu\!-\!bv\!-\!aw & \!-\!ct\!+\!du\!-\!av\!+\!bw & bt\!+\!au\!+\!dv\!+\!cw & at\!-\!bu\!-\!cv\!+\!dw \end{array} \right)_. \end{equation} It has determinant \begin{eqnarray} \det(M(\bm{q}^{}_1,\bm{q}^{}_2)) & = & (a^2+b^2+c^2+d^2)^2\cdot(t^2+u^2+v^2+w^2)^2 \nonumber \\ & = & \{(\bm{q}^{}_1)^2 \cdot (\bm{q}^{}_2)^2\}^2 \end{eqnarray} and also fulfils \begin{equation} M M^t \; = \; \sqrt{\det{M}} \cdot \mbox{\large \bf 1}^{}_4 \, . \end{equation} Consequently, when $|\bm{q}^{}_1| = |\bm{q}^{}_2| = 1$, we obtain a 4D rotation matrix and the homomorphism $M: S^3\times S^3 \longrightarrow\mbox{SO}(4)$ provides the standard double cover of the rotation group $\mbox{SO}(4)$ \cite{Koecher}, with $M(\bm{q}^{}_1,\bm{q}^{}_2) = M(-\bm{q}^{}_1,-\bm{q}^{}_2)$. The orientation reversing transformations, i.e.\ the elements of $\mbox{O}(4)\backslash \mbox{SO}(4)$, are obtained by the mapping $\bm{x} \mapsto \bm{q}^{}_1\,\overline{\bm{x}}\,\overline{\bm{q}}^{}_2$ with unit quaternions $\bm{q}^{}_1,\bm{q}^{}_2$. Let us finally note that, for non-zero quaternions, \begin{equation} R(\bm{q}^{}_1,\bm{q}^{}_2) \; := \; M(\frac{\bm{q}^{}_1}{|\bm{q}^{}_1|}, \frac{\bm{q}^{}_2}{|\bm{q}^{}_2|}) \; = \; \frac{1}{|\bm{q}^{}_1 \bm{q}^{}_2|} \, M(\bm{q}^{}_1,\bm{q}^{}_2) \end{equation} always gives a rotation matrix, which is handy for finding suitable parametrizations of groups such as $\mbox{SO}(4,\mathbb Q)$ or $\mbox{SO}(4,\mathbb Q(\tau))$ in closely related problems, see \cite{Baake} for details. \subsection*{Arguments in common} In this Section, we focus on $\mathbb I$ versus $\mathbb J$ and carry the arguments as far as possible without having to separate the two rings $\mathbb I$ and $\mathbb J$ too seriously\footnote{The case ${\cal O}=\mathbb K$ is entirely parallel to that of ${\cal O}=\mathbb I$ and need not be spelled out here. We shall mention details later when we need them.}. Thus we introduce the following notation to cover both situations simultaneously: \begin{equation} \label{defs1} \begin{array}{ccccc} {K} & \; := \; & \mathbb Q & \mbox{ or } & \mathbb Q(\mbox{\small $\sqrt{5}$}) \\ {\scriptstyle {\cal O}} & \; := \; & \mathbb Z & \mbox{ or } & \mathbb Z[\tau] \\ {\cal O} & \; := \; & \mathbb J & \mbox{ or } & \mathbb I \\ {\cal L} & \; := \; & \mathbb Z^4 & \mbox{ or } & \mathbb Z[\tau]^4 \end{array} \end{equation} By ${\cal L}$ we really mean the ring ${\scriptstyle {\cal O}} 1 + {\scriptstyle {\cal O}} \bm{i} + {\scriptstyle {\cal O}} \bm{j} + {\scriptstyle {\cal O}} \bm{k} \subset {\cal O}$, and we observe that \begin{equation} \label{inclusion} 2 {\cal O} \; \subset \; {\cal L} \; \subset \; {\cal O} \, . \end{equation} Let us first note that ${\cal O}$ is a maximal order in $\mathbb H({K})$ \cite{Vigneras}. As such, each prime ideal $\mathfrak P$ of ${\cal O}$ corresponds to one prime ideal of ${\scriptstyle {\cal O}}$, namely to $\mathfrak p := \mathfrak P\cap{\scriptstyle {\cal O}}$, and this sets up a one-to-one correspondence between their prime ideals \cite[Thm.\ 22.4]{Reiner}. {}Furthermore, we have unique factorization of each 2-sided ideal ${\mathfrak A}$ of ${\cal O}$: \begin{equation} {\mathfrak A} \; = \; \mathfrak P_1^{k_1}\cdot\ldots\cdot\mathfrak P_r^{k_r} \end{equation} where $\mathfrak P_1^{k_1},\ldots,\mathfrak P_r^{k_r}$ are prime ideals and all $k_i\geq 0$. The prime 2 (which is prime both in $\mathbb Z$ and in $\mathbb Z[\tau]$) plays a special role in this paper. In the case of $\mathbb I$, $2\mathbb I$ is the prime ideal of $\mathbb I$ lying over $2\mathbb Z[\tau]$. The case of $\mathbb J$, however, is more complicated. Here, $(1+\bm{i})\mathbb J=\mathbb J(1+\bm{i})$ is the prime ideal lying over $2\mathbb Z$ and $(1+\bm{i})^2\mathbb J=2\mathbb J$ \cite{Hurwitz}. It is this ramification of $2\mathbb Z$ in $\mathbb J$ that accounts for the stray factor in Eq.~(\ref{Hurwitz-zeta}) and, later on, in Eq.~(\ref{genfun1}). To cope with this, we shall call $\bm{a}\in\mathbb J$ an {\em odd} element of $\mathbb J$ if $\bm{a}\not\in(1+\bm{i})\mathbb J$. It is useful to note that $\bm{a}\in\mathbb J$ is odd if and only if $|\bm{a}|^2\in\mathbb Z$ (its quaternionic norm) is odd. Let $\bm{a}\in{\cal O}$. We define \begin{equation} \label{defs2} A(\bm{a}) \; := \; \{ r\in {K}\mid r\bm{a}\in{\cal O} \} \, . \end{equation} {}For $\bm{a}\neq 0$, ${\scriptstyle {\cal O}}\subset A(\bm{a})\subset {\cal O}\bm{a}^{-1} \cap {K}$. Thus $A(\bm{a})$ is a finitely generated ${\scriptstyle {\cal O}}$-module and hence is a fractional ideal of ${\scriptstyle {\cal O}}$ \cite[Ch.\ I.4]{Janusz}. It follows that $A(\bm{a})^{-1}$ is an ordinary ideal of ${\scriptstyle {\cal O}}$ and, consequently, $A(\bm{a})^{-1} = {\scriptstyle {\cal O}} c$ for some $c\in{\scriptstyle {\cal O}}$. We call $c = c(\bm{a}) \in {\scriptstyle {\cal O}}$ (which is determined up to a unit of ${\scriptstyle {\cal O}}$) the {\em content} of $\bm{a}$: $A(\bm{a}) = {\scriptstyle {\cal O}} c(\bm{a})^{-1}$. \begin{defin} We say that $\bm{a}\in{\cal O}$ is {\em ${\cal O}$-primitive} if $A(\bm{a})={\scriptstyle {\cal O}}$. \end{defin} We have just seen that $\bm{a}$ is ${\cal O}$-primitive if and only if $c(\bm{a})$ is a unit in ${\scriptstyle {\cal O}}$. So, we have \begin{lemma} \label{primitive} {}For any $\bm{a}\in{\cal O}\backslash\{0\}$, $c(\bm{a})^{-1} \bm{a}$ is a primitive element of ${\cal O}$. \hfill $\square$ \end{lemma} Let us now come to the link between submodules and similarities. \begin{defin} Let $L$ be a $\mathbb Z$-module in $\mathbb R^4$ that spans $\mathbb R^4$. ${\cal M}\subset L$ is called a {\em similarity submodule} (SSM) of $L$ if there is an $\alpha\in\mathbb R\backslash\{0\}$ and an $R\in{\it O}(4)$ so that ${\cal M}=\alpha R({\cal O})\subset L$. If $L$ is a {\em lattice}, we call ${\cal M}$ a {\em similarity sublattice} (SSL) of $L$. \end{defin} Consider the case when ${\cal M}$ is an SSM of ${\cal O}$. It is immediate that such an ${\cal M}$ is an ${\scriptstyle {\cal O}}$-submodule of ${\cal O}$. Since ${\scriptstyle {\cal O}}$ is a principal ideal domain and ${\cal O}$ is a free ${\scriptstyle {\cal O}}$-module of rank 4, we see that ${\cal M}$ is also a free ${\scriptstyle {\cal O}}$-module of rank 4. Consequently, the index $[{\cal O}:{\cal M}]$ of ${\cal M}$ in ${\cal O}$ is finite. We now come to the first crucial assertion in our classification of the similarity submodules according to their indices. \begin{prop} \label{prop1} Let ${\cal M}\subset{\cal O}$ be a similarity submodule. Then there exist $\bm{a},\bm{b}\in{\cal O}$, with $\bm{a}$ primitive, such that ${\cal M}=\bm{a}{\cal O}\bm{b}$. In addition, in the case ${\cal O}=\mathbb J$, we can arrange for $\bm{a}$ to be odd. \end{prop} {\sc Proof}: By assumption, ${\cal M}=\alpha R({\cal O})$ with $\alpha\in\mathbb R$, $\alpha\neq 0$, and $R\in\mbox{O}(4)=\mbox{O}(4,\mathbb R)$. Using (\ref{action}), and noting that $\overline{{\cal O}}={\cal O}$, we can write ${\cal M}=\bm{a}{\cal O}\bm{b}$ where $\bm{a}=(a^{}_0,a^{}_1,a^{}_2,a^{}_3)$, $\bm{b}=(b^{}_0,b^{}_1,b^{}_2,b^{}_3)$ and $\bm{a},\bm{b} \in \mathbb H(\mathbb R)$. Since $2\bm{a} \{ 1,\bm{i},\bm{j},\bm{k} \} \bm{b} \subset 2\bm{a}{\cal O}\bm{b}\subset{\cal L}$, we can infer from the explicit matrix form (\ref{matrix}) that all the matrix entries of $2 M(\bm{a},\bm{b})$ lie in ${\scriptstyle {\cal O}}$. Combining suitable entries, four at a time, we can obtain that \begin{equation} \label{proof1} 8 a^{}_i b^{}_j \in {\scriptstyle {\cal O}} \, , \quad \quad \mbox{ for all } i,j \in \{0,1,2,3\} \, . \end{equation} Since $\bm{a}{\cal O}\bm{b}=\bm{a} r {\cal O} r^{-1} \bm{b}$ for all $r\in \mathbb R\backslash\{0\}$, we can arrange that some $a^{}_i\in {K}\backslash\{0\}$, whereupon we see, via (\ref{proof1}), that we can choose $\bm{a},\bm{b}\in\mathbb H({K})$ without loss of generality. Clearing denominators and using Lemma~\ref{primitive}, we may further assume that ${\cal M}=\bm{a}{\cal O}\bm{b}$ with $\bm{a}\in{\cal O}$ and $\bm{a}$ primitive. {}From (\ref{proof1}), we now get $8b^{}_j\bm{a}\in{\cal O}$, whence $8b^{}_j\in A(\bm{a})={\scriptstyle {\cal O}}$ for all $j\in\{0,1,2,3\}$. We conclude that $8\bm{b}\in{\cal O}$. We now have to dispose of the factor 8 in order to prove the Proposition. Consider the 2-sided ideal ${\cal O}\bm{a}{\cal O}\bm{b}{\cal O}$ of ${\cal O}$. Since ${\cal O}$ is a maximal order \cite{Hurwitz,Reiner,Vigneras} in $\mathbb H({K})$, we have a unique factorization \begin{equation} \label{proof2} {\cal O}\bm{a}{\cal O}\bm{b}{\cal O} \; = \; \mathfrak P_1^{k_1}\cdot\ldots\cdot\mathfrak P_r^{k_r}\, , \quad k_1,\ldots,k_r \in \mathbb N_0 \, , \end{equation} where $\mathfrak P_1^{k_1},\ldots,\mathfrak P_r^{k_r}$ are prime ideals of ${\cal O}$ and $\mathfrak P^{}_i\cap{\scriptstyle {\cal O}}=\mathfrak p^{}_i$ are in one-to-one correspondence with distinct prime ideals of ${\scriptstyle {\cal O}}$ \cite{Reiner}. Similarly, we have ${\cal O}\bm{a}{\cal O} = \prod \mathfrak P_i^{m_i}$ and ${\cal O}\bm{b}{\cal O} = \prod \mathfrak P_i^{n_i}$, with $k_i = m_i+n_i$ for all $i\in\{1,\ldots,r\}$, $m_i\in\mathbb N_0$, $n_i\in\mathbb Z$. Since $8{\cal O}\bm{b}{\cal O}\subset{\cal O}$, the only primes for which $n_i\leq 0$ is possible are those lying over $2\in{\scriptstyle {\cal O}}$. Recall that there is exactly one prime ideal of ${\cal O}$ that corresponds to 2, and this is $\mathfrak P^{}_1=\bm{x}{\cal O}={\cal O}\bm{x}$, where $\bm{x}=(1+\bm{i})$ in the case ${\cal O}=\mathbb J$ and $\bm{x}=2$ in the case ${\cal O}=\mathbb I$. Thus we have $n_i\geq 0$ for all $i>1$. If we can now show that $n_1=0$, then $\bm{b}\in{\cal O}$ and our assertion follows. Suppose $n_1<0$. Then $m_1\geq|n_1|>0$, and we can write \begin{equation} \bm{a}{\cal O}\bm{b} \; = \; \bm{a}\bm{x}^{-|n_1|}{\cal O}\bm{x}^{|n_1|}\bm{b} \end{equation} and ${\cal O}\bm{a}\bm{x}^{-|n_1|}{\cal O}\subset\mathfrak P_1^{m_1-|n_1|} \mathfrak P_2^{m_2}\cdot\ldots\cdot\mathfrak P_r^{m_r}\subset{\cal O}$. Similarly, ${\cal O}\bm{x}^{|n_1|}\bm{b}{\cal O}\subset{\cal O}$. In the icosian case (where $\bm{x}=2\in{\scriptstyle {\cal O}}$), the primitivity of $\bm{a}$ rules out that $\bm{a}\bm{x}^{-|n_1|}\in{\cal O}$, so $n_1<0$ is impossible here. In the Hurwitzian case, since $(1+\bm{i})^2=2$ (up to units), $n_1=-1$ is still possible. Then, $\bm{a}\bm{x}^{-1},\bm{xb}\in{\cal O}$ and $\bm{a}\bm{x}^{-1}$ is still primitive. We take these as the new $\bm{a},\bm{b}$. This achieves the correct form, ${\cal M}=\bm{a}{\cal O}\bm{b}$, with $\bm{a},\bm{b}\in{\cal O}$ and $\bm{a}$ ${\cal O}$-primitive. When ${\cal O}=\mathbb J$ and $\bm{a}$ is even, we have $\bm{a}\in(1+\bm{i}){\cal O}\backslash2{\cal O}$, and we may replace $\bm{a}{\cal O}\bm{b}$ by $\bm{a}\bm{x}^{-1}{\cal O}\bm{x}\bm{b}$. \hfill $\square$ {\bf Remark 1}: Given ${\cal M}=\bm{a}^{}_1{\cal O}\bm{b}^{}_1\subset{\cal O}$ where $\bm{a}^{}_1,\bm{b}^{}_1\in\mathbb H({K})$, the above argument shows that we can constructively adjust $\bm{a}^{}_1,\bm{b}^{}_1$ to $\bm{a}=\bm{a}^{}_1 r^{-1}$, $\bm{b}=r \bm{b}^{}_1$, where $r\in{K}$ for ${\cal O}=\mathbb I$ and $r\in{K}\cup{K}(1+\bm{i})$ for ${\cal O}=\mathbb J$, so as to have ${\cal M}=\bm{a}{\cal O}\bm{b}$ with $\bm{a},\bm{b}\in{\cal O}$ and $\bm{a}$ primitive (and odd for ${\cal O}=\mathbb J$). The argument also shows that once $\bm{a}$ is adjusted to be primitive (and odd if ${\cal O}=\mathbb J$) then $\bm{b}$ necessarily lies in ${\cal O}$. Since we are ultimately interested in the similarity submodules, and not so much in the actual self-similarities themselves, we have to draw the attention now to the symmetries of our maximal orders. The group of units ${\cal O}^{\times}$ of ${\cal O}$ is, geometrically, a finite root system $\Delta$ of type\footnote{We are using the symbol $D_4$ both for the root system and for the corresponding root lattice. The convex hull of the 24 roots of $D_4$ is the regular 24-cell \cite[p.\ 292]{Cox73}, mentioned in an earlier footnote. RVM would like to take this opportunity to note that in \cite{CMP} the root system $D_4$ was inexplicably left out of the classification of subroot systems of the root system of type $H_4$. All such $D_4$ subroot systems are conjugate by the Weyl group of $H_4$.} $D_4$ (resp.\ $H_4$) for $\mathbb J$ (resp.\ $\mathbb I$). Any self-similarity of ${\cal O}$ that is {\em surjective} is necessarily an isometry (norm preserving) and so must map $\Delta$ onto itself. Conversely, any isometry which stabilizes $\Delta$ will also stabilize its ${\scriptstyle {\cal O}}$-span which is ${\cal O}$. Thus \begin{equation} \label{stab} \mbox{stab}^{}_{{\rm O}(4)} {\cal O} \; = \; \mbox{Aut}(\Delta) \, . \end{equation} {}For ${\cal O}=\mathbb I$, $\mbox{Aut}(\Delta)$ is the Weyl group $W(\Delta)$ of $\Delta$, which is $H_4$, and consequently all elements of $\mbox{Aut}(\Delta)^{+}$ (the orientation preserving part of $\mbox{Aut}(\Delta)$) are realized by mappings $M(\bm{u},\bm{v})$ with $\bm{u},\bm{v} \in \mathbb I^{\times}$, i.e.\ units. {}For ${\cal O}=\mathbb J$, $[\mbox{Aut}(\Delta):W(\Delta)]=6$, the additional symmetry being due to the diagram automorphisms of $D_4$. In fact, $\mbox{Aut}(\Delta)$ is the Weyl group of $F_4$, and the root system of type $F_4$ can be realized explicitly as \begin{equation} \tilde{\Delta} \; := \; \Delta \cup \{ \pm\bm{u}\pm\bm{v}\mid\bm{u},\bm{v}\in \{1,\bm{i},\bm{j},\bm{k}\}, \; \bm{u}\neq\bm{v} \} \, , \end{equation} i.e.\ by adjoining to $\Delta$ the elements of $\mathbb J$ of square length 2. This time, $\mbox{Aut}(\Delta)^{+}$ is realized as the set of mappings $M(\bm{u},\bm{v})$ with either $\bm{u},\bm{v}\in\Delta$ or $\bm{u},\bm{v}\in(\tilde{\Delta}\backslash\Delta)/\sqrt{2}$. It will be observed that all the elements of $\tilde{\Delta}\backslash\Delta$ lie in the ideal $(1+\bm{i})\mathbb J$ (in fact they are all the generators of this ideal). So, we have proved \begin{prop} \label{symmetries} The orientation preserving self-similarities of ${\cal O}$ onto itself are precisely the maps $M(\bm{u},\bm{v})$ for $\bm{u},\bm{v}\in{\cal O}$ with $|\bm{u}|=|\bm{v}|=1$ and, in the case ${\cal O}=\mathbb J$, also the maps ${1\over 2}M(\bm{u},\bm{v})$ for $\bm{u},\bm{v}\in{\cal O}$ with $|\bm{u}|^2=|\bm{v}|^2=2$. \hfill $\square$ \end{prop} We say that an SSM $\bm{a}{\cal O}\bm{b}$ is given in {\em canonical form} if $\bm{a},\bm{b}\in{\cal O}$ with $\bm{a}$ being ${\cal O}$-primitive (and with $\bm{a}$ odd if ${\cal O}=\mathbb J$). \begin{prop} \label{comp} Similarity submodules $\bm{a}^{}_1 {\cal O} \bm{b}^{}_1$ and $\bm{a}^{}_2 {\cal O} \bm{b}^{}_2$ written in canonical form are equal if and only if both $\bm{a}_1^{-1}\bm{a}^{}_2$ and $\bm{b}_2^{}\bm{b}^{-1}_1$ are units in ${\cal O}$. \end{prop} {\sc Proof}: Suppose that $\bm{a}^{}_1 {\cal O} \bm{b}^{}_1=\bm{a}^{}_2 {\cal O} \bm{b}^{}_2$. Then $\bm{a}^{-1}_1\bm{a}^{}_2{\cal O}\bm{b}^{}_2\bm{b}^{-1}_1={\cal O}$. According to Prop.~\ref{symmetries}, one of two things may happen: \newline (i) There is an $r\in{K}$ so that $r\bm{a}^{-1}_1\bm{a}^{}_2=\bm{u}\in{\cal O}^{\times}$ (and $r^{-1}\bm{b}^{}_2\bm{b}^{-1}_1=\bm{v}\in{\cal O}^{\times}$). Then, $r\bm{a}^{}_2=\bm{a}^{}_1\bm{u}\in{\cal O}$ gives $r\in{\scriptstyle {\cal O}}$ since $\bm{a}^{}_2$ is primitive. Likewise, $r^{-1}\bm{a}^{}_1=\bm{a}^{}_2\bm{u}^{-1}_{}\in{\cal O}$ gives $r^{-1}\in{\scriptstyle {\cal O}}$, so $r\in{\scriptstyle {\cal O}}^{\times}\subset{\cal O}^{\times}$ and we are done in this case. \newline (ii) We are in the case ${\cal O}=\mathbb J$ and there is an $r\in{K}$ so that $2r\bm{a}^{-1}_1\bm{a}^{}_2=(1+\bm{i})\bm{u}$, $\bm{u}\in{\cal O}^{\times}$. This gives $2r\bm{a}^{}_2\in{\cal O}$ and $r^{-1}\bm{a}^{}_1=\bm{a}^{}_2\bm{u}_{}^{-1}(1-\bm{i}) \in{\cal O}$ whence $2r,r^{-1}\in\mathbb Z$. Thus $r=\pm1,\pm{1\over 2}$. Since $\bm{a}^{}_1$ and $\bm{a}^{}_2$ are both odd and $2\in(1+\bm{i})^2 {\cal O}^{\times}$, none of these values of $r$ is possible. \newline The reverse direction is clear. \hfill $\square$ We are now in the position to formulate the main result of this section. \begin{theorem} \label{commonthm} The number of similarity submodules of ${\cal O}$ of a given index is a multiplicative arithmetic function. Its Dirichlet series generating function is given by \begin{equation} {}F^{}_{{\cal O}}(s) \; = \; \frac{(\zeta^{}_{{\cal O}}(s))^2} {\zeta^{}_{{K}}(4s)} \cdot \cases{\frac{1}{1+4^{-s}} , & if ${\cal O}=\mathbb J$ \cr 1 , & if ${\cal O}=\mathbb I$ \, . } \end{equation} \end{theorem} {\sc Proof}: As a result of Prop.~\ref{symmetries} and Prop.~\ref{comp}, any SSM ${\cal M}$ of ${\cal O}$ can be uniquely written as \begin{equation} \label{fac} {\cal M} \; = \; \bm{a} {\cal O} {\cal O} \bm{b} \, , \end{equation} i.e.\ as a product of a right and a left ideal of ${\cal O}$, where $\bm{a}$ is ${\cal O}$-primitive (and also odd if ${\cal O}=\mathbb J$). Any right ideal $\bm{d}{\cal O}$ can be written {\em uniquely} as a product $c(\bm{d}) {\scriptstyle {\cal O}} \bm{a} {\cal O}$, where $c(\bm{d})$ (the content of $\bm{d}$) is in ${\scriptstyle {\cal O}}$ and $\bm{a}\in{\cal O}$ is ${\cal O}$-primitive. In addition, we have the formula $[{\cal O} : \bm{d}{\cal O}] = [{\scriptstyle {\cal O}}:c(\bm{d}){\scriptstyle {\cal O}}]^4 \cdot [{\cal O}:\bm{a}{\cal O}]$. Thus the Dirichlet series for the {\em primitive} right ideals of ${\cal O}$ (those $\bm{a}{\cal O}$ with $\bm{a}$ primitive) is the quotient of two zeta functions, \begin{equation} \label{prim} \zeta^{}_{{\cal O}}(s)/\zeta^{}_{{K}}(4s) \, . \end{equation} In the case ${\cal O}=\mathbb I$, the factorization (\ref{fac}) leads at once to $F^{}_{\mathbb I}(s)=(\zeta^{}_{\mathbb I}(s))^2/\zeta^{}_{\mathbb Q(\tau)}(4s)$. In the case ${\cal O}=\mathbb J$, writing (\ref{prim}) explicitly as an Euler product, we find that the contribution for the prime 2 is $(1+4^{-s})$. This corresponds to the fact that the primitive right ideals are either of the form $\bm{a}\mathbb J$ with $\bm{a}$ odd or of the form $\bm{a}(1+\bm{i})\mathbb J$ with $\bm{a}$ odd. The latter ones are to be removed from the counting (because $(1+\bm{i})\mathbb J=\mathbb J(1+\bm{i})$ otherwise leads to doubly counting them), and the corresponding Dirichlet series is obtained by removing the factor $(1+4^{-s})$. This gives the result claimed. \hfill $\square$ \subsection*{Results: lattices} Let us now consider the Hurwitzian case ${\cal O}=\mathbb J$ in detail, and also the lattice ${\cal L}=\mathbb Z^4$. If $\bm{x}=(1+\bm{i})$, we have $\bm{x}{\cal O}={\cal O}\bm{x}$ and also $\bm{x}{\cal L}={\cal L}\bm{x}$. These lattices are related by \begin{equation} \label{subset1} \bm{x}{\cal O} \; \stackrel{2}{\subset} \; {\cal L} \; \stackrel{2}{\subset} \; {\cal O} \end{equation} where the integer on top of the inclusion symbol is the corresponding index. \begin{lemma} \label{back} If $\bm{a}{\cal L}\bm{b}\subset{\cal L}$, then there exist $\bm{a}^{}_1,\bm{b}^{}_1 \in{\cal O}$, with $\bm{a}^{}_1$ odd and ${\cal O}$-primitive, such that $\bm{a}{\cal L}\bm{b}=\bm{a}^{}_1{\cal L}\bm{b}^{}_1$. \end{lemma} {\sc Proof}: Let $\bm{a}{\cal L}\bm{b}\subset{\cal L}$. We can assume, without loss of generality, that $\bm{a},\bm{b}\in\mathbb H(\mathbb Q)$ and, by using a suitable scaling and the fact that $\bm{x}{\cal L}={\cal L}\bm{x}$, we may even assume that $\bm{a}\in{\cal O}$ and that $\bm{a}$ is ${\cal O}$-primitive and odd. Since $\bm{a}{\cal O}\bm{x}\bm{b}\subset\bm{a}{\cal L}\bm{b}\subset{\cal O}$, the conditions on $\bm{a}$ already show that $\bm{x}\bm{b}\in{\cal O}$ (see Remark 1). Write $\bm{b}=\bm{x}^{-1} \bm{c}$ with $\bm{c}\in{\cal O}$. Consider $[{\cal L}:\bm{a}{\cal L}\bm{b}]= |\bm{a}|^4 \, |\bm{b}|^4 = \frac{|\bm{a}|^4 \, |\bm{c}|^4}{4} \in\mathbb Z$. Since $\bm{a}$ is odd, it follows that $4 \mid |\bm{c}|^2$ and hence $\bm{x}\mid\bm{c}$, because any even element in $\mathbb J$ is of the form $(1+\bm{i})^r$ times an odd element \cite{Hurwitz}. Consequently, $\bm{b}\in{\cal O}$ and we conclude that $\bm{a}{\cal L}\bm{b}\subset{\cal L}$ implies that we can rearrange the quaternions in the way claimed. \hfill $\square$ This provides a link between the SSL problems for ${\cal L}$ and for ${\cal O}$. The difference between the counting arises as follows. The symmetry group of ${\cal O}$ is isomorphic with the Weyl group of $F_4$ (see above), while that of ${\cal L}$, which is the Weyl group of $B_4$, is a subgroup of index 3. As we shall show, this only influences the number of SSLs of even index when going from ${\cal O}$ to ${\cal L}$. \begin{theorem} \label{thm-lattices} The possible indices of similarity sublattices of hypercubic lattices in 4D are precisely the squares of rational integers. The number of SSLs of given index is a multiplicative arithmetic function. {}For the case of $\mathbb J=D_4^*$, the corresponding Dirichlet series generating function $F^{}_{\mathbb J}$ reads \begin{equation} \label{genfun1} {}F^{}_{\mathbb J}(s) \; = \; \frac{(\zeta^{}_{\mathbb J}(s))^2}{(1+4^{-s})\,\zeta(4s)} \; = \; \frac{(1-2^{1-2s})^2}{1+4^{-s}} \cdot \frac{(\zeta(2s)\zeta(2s-1))^2}{\zeta(4s)} \, . \end{equation} The same series also applies to the lattice $D_4$, while for the primitive hypercubic lattice, $\mathbb Z^4$, it reads \begin{equation} \label{genfun2} {}F^{}_{\mathbb Z^4}(s) \; = \; (1+\frac{2}{4_{}^s}) \cdot {}F^{}_{\mathbb J}(s) \, . \end{equation} \end{theorem} {\sc Proof}: The statement about $F^{}_{\mathbb J}(s)$ follows directly from Theorem \ref{commonthm} and from Eq.~(\ref{Hurwitz-zeta}). It is rather easy to see from the Euler product representation that precisely all squares of integers occur as indices. In order to extend this to ${\cal L}$, we have to understand how the different symmetries lead to different countings. Assume $\bm{a}{\cal L}\bm{b}\subset{\cal L}$. Due to Lemma~\ref{back}, we may assume that $\bm{a},\bm{b}\in{\cal O}$ with $\bm{a}$ odd and ${\cal O}$-primitive, i.e.\ we assume canonical form. By Prop.~\ref{comp}, $\bm{a}{\cal O}\bm{b}=\bm{a}^{}_1{\cal O}\bm{b}^{}_1$ if and only if there are units $\bm{u},\bm{v}\in{\cal O}^{\times}$ with $\bm{a}^{}_1=\bm{a}\bm{u}$ and $\bm{b}^{}_1=\bm{v}\bm{b}$. However, the unit group of ${\cal L}$ is only the quaternion group, $Q=\{\pm 1,\pm\bm{i},\pm\bm{j},\pm\bm{k}\}$, and $Q$ is a normal subgroup of ${\cal O}^{\times}$ with ${\cal O}^{\times}/Q\simeq\mathbb Z/3\mathbb Z$. We may take $\bm{t}=(1,1,1,1)/2$ as a suitable representative of a generator of this cyclic group in ${\cal O}^{\times}$ \cite[{\S} 26]{duval}. Note that, for $\bm{u},\bm{v}\in{\cal O}^{\times}$, we have $\bm{a}\bm{u}{\cal L}\bm{v}\bm{b}=\bm{a}{\cal L}\bm{b}$ if and only if $\bm{uv}\in Q$. So the single SSL $\bm{a}{\cal O}\bm{b}$ of ${\cal O}$ {\em may} give rise to three different SSLs of ${\cal L}$, namely to $\bm{a}{\cal L}\bm{b}$, $\bm{a}{\cal L}\bm{t}\bm{b}$, $\bm{a}{\cal L}\bm{t}^2\bm{b}$. Whether or not this happens depends on whether or not the latter two are actually in ${\cal L}$. Now, if $\bm{a}{\cal L}\bm{b}$ and $\bm{a}{\cal L}\bm{t}\bm{b}$ are both in ${\cal L}$, then so is $\bm{a}({\cal L}+{\cal L}\bm{t})\bm{b}=\bm{a}{\cal O}\bm{b}$, whence also $\bm{a}{\cal L}\bm{t}^2\bm{b}\subset{\cal L}$. Similarly, $\bm{a}{\cal L}\bm{t}^2\bm{b}\subset{\cal L}$ implies $\bm{a}{\cal L}\bm{t}\bm{b}\subset{\cal L}$, too. Thus $\bm{a}{\cal O}\bm{b}\subset{\cal O}$ gives rise to 3 different SSLs of ${\cal L}$ if and only if $\bm{a}{\cal O}\bm{b}\subset{\cal L}$. However, $\bm{a}{\cal O}\bm{b}\subset{\cal L}$ implies that $[{\cal O}:\bm{a}{\cal O}\bm{b}]$ is even because $[{\cal O}:{\cal L}]=2$. Conversely, $[{\cal O}:\bm{a}{\cal O}\bm{b}]$ even implies that $|\bm{a}|^4\,|\bm{b}|^4$ is divisible by 4, so $\bm{a}$ or $\bm{b}$ must be even and hence $\bm{a}{\cal O}\bm{b}\subset\bm{x}{\cal O}\subset{\cal L}$. In short, the 3 SSL situation occurs if and only if $[{\cal O}:\bm{a}{\cal O}\bm{b}]$ is even. As a consequence, the counting function for ${\cal L}$ is still multiplicative, and the modification in the Euler product expansion occurs only in the factor that belongs to the prime 2. It is easy to check that the result is that given in the Theorem. \hfill $\square$ If we take into account that the possible indices are always squares, it is reasonable to define the appropriate coefficients as follows, \begin{equation} \label{newseries} {}F^{}_{{\cal L}}(s) \; = \; \sum_{m=1}^{\infty} \frac{f^{}_{{\cal L}}(m)}{m^{2s}} \, . \end{equation} So, the coefficients actually are \begin{equation} f^{}_{{\cal L}}(m) \; = \; |\{{\cal L}' \mbox{ is SSL of } {\cal L} \mid [{\cal L} : {\cal L}'] = m^2 \} | \, . \end{equation} To simplify explicit formulas here and later on, we introduce the function \begin{equation} \label{help} g(n,r) \; := \; (r+1) n^r + 2 \frac{1-(r+1) n^r + r\, n^{r+1}} {(n-1)^2} \end{equation} for integers $r\geq 0$ and $n>1$. Note that $g(n,0)=1$. An explicit expansion of the Euler factors now gives the following result. \begin{coro} The arithmetic function $f^{}_{\mathbb J}(m)$ is multiplicative. It is given by \begin{equation} \label{afun1} f^{}_{\mathbb J}(p^r) \; = \; \cases{ 1 , & if $p=2$ \cr g(p,r), & if $p$ is an odd prime } \end{equation} where $r\geq 0$ in all cases. Similarly, $f^{}_{\mathbb Z^4}(m)$ is a multiplicative arithmetic function. It is related to $f^{}_{\mathbb J}(m)$ via \begin{equation} \label{modif} f^{}_{\mathbb Z^4}(m) \; = \; \cases{f^{}_{\mathbb J}(m) , & $m$ odd \cr 3 \cdot f^{}_{\mathbb J}(m) , & $m$ even.} \end{equation} \end{coro} The first few terms of $F^{}_{\mathbb J}(s)$ read explicitly \begin{equation} {}F^{}_{\mathbb J}(s) \; = \; \mbox{\small $ 1+\frac{1}{4^s}+\frac{8}{9^s}+\frac{1}{16^s}+ \frac{12}{25^s}+\frac{8}{36^s}+\frac{16}{49^s}+ \frac{1}{64^s}+\frac{41}{81^s}+\frac{12}{100^s}+ \frac{24}{121^s}+\frac{8}{144^s}+ \cdots $ } \end{equation} while those for $F^{}_{\mathbb Z^4}(s)$ follow easily from (\ref{modif}). They are now listed as \cite[sequence A 045771]{Sloane} and \cite[sequence A 035292]{Sloane}, respectively. Note that Eq.~(\ref{afun1}) implies that the SSMs of $\mathbb J$ of index $4^r$ are unique -- they are, in fact, just the 2-sided ideals $(1+\bm{i})^r \mathbb J$. Let us now, in line with the previous examples, briefly consider the asymptotic behaviour of the coefficients. Since $\zeta(s)\neq 0$ in $\{\mbox{Re}(s)\geq 1\}$, it is clear that $F^{}_{{\cal L}}(s)$ is meromorphic in the same half-plane, with only one pole which is of second order and located at $s=1$. This is true both of ${\cal L}=\mathbb J$ and ${\cal L}=\mathbb Z^4$. Using the Dirichlet series (\ref{newseries}) and applying the results from the Appendix to $F^{}_{{\cal L}}(s/2)$, we get the following \begin{coro} The coefficients $f^{}_{{\cal L}}(m)$ grow faster than linear on average for large $m$, and we have the asymptotic behaviour \begin{equation} \sum_{m\leq x} f^{}_{{\cal L}}(m) \; \simeq \; C^{}_{{\cal L}} \cdot x^2 \log(x) \quad \quad \quad \mbox{(as $x\to\infty$)} \end{equation} where the constant is given by \begin{equation} C^{}_{{\cal L}} \; = \; \frac{1}{2} \, \mbox{\rm res}_{s=1} \, ((s-1) {}F^{}_{{\cal L}}(s)) \; = \; \frac{1}{4} \cdot \cases{ 1 \, , & if ${\cal L}=\mathbb J$ \cr \frac{3}{2} \, , & if ${\cal L}=\mathbb Z^4$ \, . } \end{equation} \end{coro} Note that this really is an asymptotic result, and that, numerically, the estimates of $C^{}_{{\cal L}}$ converge rather slowly (from above) to the values given in the Corollary. \subsection*{Results: modules} Let us first state the result for the icosian ring $\mathbb I$ itself. \begin{theorem} \label{thm-modules} The possible indices of similarity submodules of the icosian ring $\mathbb I$ are the squares of rational integers that can be represented by the quadratic form $x^2 + xy - y^2$. The number of SSMs of given index is a multiplicative arithmetic function, its Dirichlet series generating function reads \begin{equation} \label{genfun3} {}F^{}_{\mathbb I}(s) \; = \; \frac{(\zeta^{}_{\mathbb I}(s))^2} {\zeta^{}_{{K}}(4s)} \; = \; \frac{(\zeta^{}_{{K}}(2s)\zeta^{}_{{K}}(2s-1))^2} {\zeta^{}_{{K}}(4s)} \end{equation} where ${K}=\mathbb Q(\tau)$. \end{theorem} The proof follows immediately from Theorem \ref{commonthm} in combination with Eq.~(\ref{icosian-zeta}). Note that the possible indices are just the squares of the possible norms of ideals in $\mathbb Q(\tau)$ and hence of the form given. Taking this into account, we write \begin{equation} {}F^{}_{\mathbb I}(s) \; = \; \sum_{m=1}^{\infty} \frac{f^{}_{\mathbb I}(m)}{m^{2s}} \end{equation} in analogy to above, and obtain, by an explicit expansion of the Euler factors, the following result (compare \cite[sequence A 035284]{Sloane}). \begin{coro} The arithmetic function $f^{}_{\mathbb I}(m)$ is multiplicative and given by \begin{equation} f^{}_{\mathbb I}(p^r) \; = \; \cases{ g(5,r) , & if $p=5$ \cr 0 , & if $p\equiv\pm 2$ mod 5 and $r$ is odd \cr g(p^2,\ell), & if $p\equiv\pm 2$ mod 5 and $r=2\ell$ \cr \sum_{s=0}^{r} g(p,s) g(p,r\!-\!s) , & if $p\equiv\pm 1$ mod 5 } \end{equation} where always $r\geq 0$ and $g$ is the function defined in Eq.~(\ref{help}). \end{coro} The first few terms of $F^{}_{\mathbb I}(s)$ read explicitly \begin{equation} {}F^{}_{\mathbb I}(s) \; = \; \mbox{\small $ 1+\frac{10}{16^s}+\frac{12}{25^s}+\frac{20}{81^s}+ \frac{48}{121^s}+\frac{66}{256^s}+\frac{80}{361^s}+ \frac{120}{400^s}+\frac{97}{625^s}+\frac{120}{841^s}+ \frac{128}{961^s}+ \cdots $ } \end{equation} Let us briefly look at the $10$ SSMs of $\mathbb I$ of index $16$. {}From Eq.~(\ref{ico-zeta-2}) it is obvious that they are just the 5 left ideals $\mathbb I\bm{a}$ and the 5 right ideals $\bm{a}\mathbb I$, with suitable generators $\bm{a}$ with $\mbox{N}(|\bm{a}|^2)=16$. Note that none of them is 2-sided. {}Finally, we can again determine the asymptotic behaviour along the lines used before. $F^{}_{\mathbb I}(s/2)$ is holomorphic in the half-plane $\{\mbox{Re}(s)\geq 2 \}$, with a single second-order pole at $s=2$. With the results from the Appendix, we then obtain \begin{coro} {}For $x\to\infty$, the asymptotic behaviour of the coefficients $f^{}_{\mathbb I}(m)$ is \begin{equation} \sum_{m\leq x} f^{}_{\mathbb I}(m) \; \sim \; \frac{6 (\log(\tau))^2}{5\sqrt{5}} \, x^2 \log(x) \; \simeq \; 0.124271 \, x^2 \log(x) \, . \end{equation} \end{coro} At this point, it would be interesting to relate these findings to the corresponding ones for the $\mathbb Z[\tau]$-modules ${\cal L}=\mathbb Z[\tau]^4$ and ${\cal M}=\mathbb J[\tau]=\mathbb I\cap\mathbb I'$. Since this requires a lot more effort than in the previous case ($\mathbb Z^4$ versus $\mathbb J$), we postpone it, and rather state the result for the cubian\footnote{The term ``octonian'' would be more natural a choice, but it has already been taken!} maximal order $\mathbb K$ in $\mathbb H(\mathbb Q(\mbox{\small $\sqrt{2}$}))$. \begin{theorem} \label{thm-modules-2} The possible indices of similarity submodules of the cubian ring $\mathbb K$ are the squares of rational integers that can be represented by the quadratic form $x^2 - 2 y^2$. The number of SSMs of given index is a multiplicative arithmetic function, its Dirichlet series generating function reads \begin{eqnarray} \label{genfun4} {}F^{}_{\mathbb K}(s) & = & \frac{(\zeta^{}_{\mathbb K}(s))^2} {\zeta^{}_{{K}}(4s)} \; = \; \frac{(\zeta^{}_{{K}}(2s)\zeta^{}_{{K}}(2s-1))^2} {\zeta^{}_{{K}}(4s)} \\ & = & \mbox{\small $ 1+\frac{6}{4^s}+\frac{22}{16^s}+\frac{32}{49^s}+ \frac{66}{64^s}+\frac{20}{81^s}+\frac{192}{196^s}+ \frac{178}{256^s}+\frac{72}{289^s}+\frac{120}{324^s}+ \frac{96}{529^s}+{52\over625^s}+ \cdots $ } \nonumber \end{eqnarray} where ${K}=\mathbb Q(\mbox{\small $\sqrt{2}$})$. \end{theorem} The proof follows directly from the proof of Theorem \ref{commonthm}, since literally every step taken for the icosian ring translates into one here, with ${K}=\mathbb Q(\mbox{\small $\sqrt{2}$})$, ${\scriptstyle {\cal O}}=\mathbb Z[\mbox{\small $\sqrt{2}$}]$, ${\cal O}=\mathbb K$, and ${\cal L}=\mathbb Z[\mbox{\small $\sqrt{2}$}]^4$. Note that now, since 2 is not a prime in $\mathbb Z[\mbox{\small $\sqrt{2}$}]$ (it actually ramifies there), $\mathfrak P^{}_1=\mbox{\small $\sqrt{2}$}\,\mathbb K=(2+\mbox{\small $\sqrt{2}$})\mathbb K$ is the prime ideal of $\mathbb K$ sitting on top of the prime ideal $(2+\mbox{\small $\sqrt{2}$})\mathbb Z[\mbox{\small $\sqrt{2}$}]$, and the arguments in the proofs have to be adjusted accordingly. One can again work out the coefficients $f^{}_{\mathbb K}(m)$ explicitly (see \cite[sequence A 035285]{Sloane}) \begin{equation} f^{}_{\mathbb K}(p^r) \; = \; \cases{ g(2,r) , & if $p=2$ \cr 0 , & if $p\equiv\pm 3$ (8) and $r$ is odd \cr g(p^2,\ell), & if $p\equiv\pm 3$ (8) and $r=2\ell$ \cr \sum_{s=0}^{r} g(p,s) g(p,r\!-\!s) , & if $p\equiv\pm 1$ (8) } \end{equation} and the asymptotic behaviour, using the Appendix, is \begin{equation} \sum_{m\leq x} f^{}_{\mathbb K}(m) \; \sim \; \frac{15 (\log(1+\mbox{\small $\sqrt{2}$}))^2}{22\sqrt{2}} \, x^2 \log(x) \; \simeq \; 0.374519 \, x^2 \log(x) \, . \end{equation} \subsection*{Concluding remarks} As we have demonstrated above, the similarity submodules of certain 4D $\mathbb Z$-modules can be classified by means of algebraic methods based on quaternionic algebras and their maximal orders. Together with the results of \cite{BM1,BM2}, this essentially covers the cases related to root systems in dimensions $d\leq 4$. Although we did not emphasize it, one can also determine the actual semigroups of self-similarities of these modules explicitly, notably through the canonical representation of SSMs (Prop.\ \ref{symmetries}) and their uniqueness up to symmetries (Prop.\ \ref{comp}). We have described this in more detail for other cases \cite{BM1}, and the interested reader will find no difficulty to extend that approach to this situation. One application is concerned with the symmetries of coloured versions of the lattices and modules under consideration. Assume that $L$ has a non-trivial (irreducible) point symmetry, and a sublattice $L'$ which is the image of a self-similarity of $L$ of index $m=[L:L']>1$. If we assign $m$ different colours to the cosets of $L'$, certain subgroups of the point group of $L'$ (which is conjugate to that of $L$) will give rise to a {\em colour symmetry} in the sense that their elements induce a unique, global permutation of the colours, compare \cite{Schwarz2,Ron} and references therein. This is also closely related to the classification of coincidence site submodules, i.e.\ of submodules that can be written as the intersection of the original module with a rotated copy of itself, see \cite{Baake} for background and some recent results. Here are several open questions, particularly in spaces of even dimension, which the above results should help to solve for dimension four. {}Finally, one would like to know to what extent a generalization of our results is possible. The root lattices seem to form a sufficiently well-behaved class of objects to try, and some partial answers on the existence of similarity sublattices and their possible indices are given in \cite{CRS}. We are, however, not aware of general results along the lines discussed here, i.e.\ including the determination of the number of SSLs of a given index, nor even of a method to overcome the dependence on special features such as the arithmetic of quaternions. \vspace{5mm} \subsection*{Acknowledgements} We are grateful to Alfred Weiss for his help in understanding the arithmetic of quaternionic maximal orders. It is our pleasure to thank Peter Pleasants and Johannes Roth for several helpful discussions and Neil Sloane for communication of material prior to publication. This work was supported by the German Science {}Foundation (DFG) and by the Natural Sciences and Engineering Research Council of Canada (NSERC). \clearpage \subsection*{Appendix} In what follows, we briefly summarize the results from analytic number theory that we need to determine certain asymptotic properties of the coefficients of Dirichlet series generating functions. {}For the general background, we refer to \cite{Apostol} and \cite{Zagier}. Consider a Dirichlet series of the form $F(s)=\sum_{m=1}^{\infty} a(m) m^{-s}$. We are mainly interested in the quantity $A(x)=\sum_{m\leq x} a(m)$ and its behaviour for large $x$. Let us give one classical result (based upon Tauberian theorems) for the case that $a(m)$ is real and non-negative. \begin{theorem} \label{meanvaluetheorem} Let $F(s)$ be a Dirichlet series with non-negative coefficients which converges for ${\rm Re}(s) > \alpha > 0$. Suppose that $F(s)$ is holomorphic at all points of the line $\{ {\rm Re}(s) = \alpha \}$ except at $s=\alpha$. Here, when approaching $\alpha$ from the half-plane right of it, we assume $F(s)$ to have a singularity of the form $F(s) = g(s) + h(s)/(s-\alpha)^{n+1}$ where $n$ is a non-negative integer, and both $g(s)$ and $h(s)$ are holomorphic at $s=\alpha$. Then we have, as $x\rightarrow\infty$, \begin{equation} \label{meanvalues} A(x) \; := \; \sum_{m\leq x} a(m) \; \sim \; \frac{h(\alpha)}{\alpha\cdot n!} \; x_{}^{\alpha} \, (\log(x))_{}^n \, . \end{equation} \end{theorem} The proof follows easily from Delange's theorem, e.g.\ by taking $q=0$ and $\omega=n$ in Tenenbaum's formulation of it, see \cite[ch.\ II.7, Thm.\ 15]{Tenenbaum} and references given there. Note that Delange's theorem is a lot more general in that is still gives results for other local behaviour of $F(s)$ in the neighbourhood of $s=\alpha$, in particular for $n$ not an integer and even for combinations with logarithmic singularities. Let us also point out that there are various extensions to Dirichlet series with complex coefficients, e.g.\ Thm.~1 on p.\ 311 of \cite{Lang}, and even stronger results (with good error estimates) for multiplicative arithmetic functions $a(m)$ with values in the unit disc, see \cite[ch.\ I, {\S} 3.8 and ch.\ III, {\S} 4.3]{Tenenbaum} for details. The critical assumption in Theorem~\ref{meanvaluetheorem} is the behaviour of $F(s)$ along the entire line $\{ \mbox{Re}(s) = \alpha \}$. In all cases that appear in this article, this can be checked explicitly. To do so, we have to know a few properties of the Riemann zeta function, $\zeta(s)$, and of the Dedekind zeta functions of $\mathbb Q(\tau)$ and $\mathbb Q(\mbox{\small $\sqrt{2}$})$. It is well known that $\zeta(s)$ is a meromorphic function in the complex plane, and that is has a sole simple pole at $s=1$ with residue 1 \cite[Thm.\ 12.5(a)]{Apostol}. It has no zeros in the half-plane $\{ \mbox{Re}(s)\geq 1 \}$ \cite[ch.\ II.3, Thm.\ 9]{Tenenbaum}. The values of $\zeta(s)$ at positive even integers are known \cite[Thm.\ 12.17]{Apostol} and we have \begin{equation} \zeta(2) \; = \; \frac{\pi^2}{6} \quad , \quad \zeta(4) \; = \; \frac{\pi^4}{90} \, . \end{equation} This is all we need to know for this case. The Dedekind zeta function of ${K}=\mathbb Q(\tau)$ has some similarly nice properties. It follows from \cite[Thm.\ 4.3]{Wash} or from \cite[\S 11, Eq.\ (10)]{Zagier} that it can be written as \begin{equation} \zeta^{}_{\mathbb Q(\tau)}(s) \; = \; \zeta(s) \cdot L(s,\chi) \end{equation} where $L(s,\chi)$ is the $L$-series of the primitive Dirichlet character \cite[Ch.\ 6.8]{Apostol} $\chi$ defined by \begin{equation} \chi(n) \; = \; \cases{0, & $n\equiv 0$ (5) \cr 1, & $n\equiv\pm 1$ (5) \cr -1, & $n\equiv\pm 2$ (5) \, . } \end{equation} Since $\chi$ is not the principal character, $L(s,\chi)=\sum_{m=1}^{\infty}\chi(m)\, m^{-s}$ is an entire function \cite[Thm.\ 12.5]{Apostol}. Consequently, $\zeta^{}_{\mathbb Q(\tau)}(s)$ is meromorphic, and its only pole is simple and located at $s=1$. The residue is $L(1,\chi)$ and from \cite[Thm.\ 4.9]{Wash} we get \begin{equation} \mbox{res}_{s=1} \, \zeta^{}_{\mathbb Q(\tau)}(s) \; = \; L(1,\chi) \; = \; \frac{2\log(\tau)}{\sqrt{5}} \; \simeq \; 0.430409 \, . \end{equation} Since $L(s,\chi)\neq 0$ for $\mbox{Re}(s)>1$, see \cite[p.\ 31]{Wash}, $\zeta^{}_{\mathbb Q(\tau)}(s)$ cannot vanish there either. Also, one can again calculate the values of $\zeta^{}_{\mathbb Q(\tau)}(s)$ at positive even integers. This is done by means of the functional equation of $L(s,\chi)$ \cite[p.\ 30]{Wash} and the knowledge of the values of $L$-functions at negative integers in terms of generalized Bernoulli numbers \cite[Thm.\ 4.2]{Wash}, see \cite[Prop.\ 4.1]{Wash} for a formula for them. Working this out explictly for $s=2$ and $s=4$ gives \begin{equation} \zeta^{}_{\mathbb Q(\tau)}(2) \; = \; \frac{2\pi^4}{75 \sqrt{5}} \quad , \quad \zeta^{}_{\mathbb Q(\tau)}(4) \; = \; \frac{4\pi^8}{16875 \sqrt{5}} \, . \end{equation} Let us add that these results, and those to follow, can also be found, in rather explicit form, in \S 9 and \S 11 of \cite{Zagier}. In the same way, one can determine the zeta function of $\mathbb Q(\mbox{\small $\sqrt{2}$})$ and its properties. One has $\zeta^{}_{\mathbb Q({\scriptscriptstyle \sqrt{2}})}(s) = \zeta(s) L(s,\chi)$, now with the primitive Dirichlet character $\chi(n)=0,1,-1$ for $n$ even, $n\equiv\pm 1$ (8), $n\equiv\pm 3$ (8), respectively. This zeta function has again only one simple pole, at $s=1$, with residue $L(1,\chi)=\log(1+\mbox{\small $\sqrt{2}$})/\mbox{\small $\sqrt{2}$}\simeq 0.623225$. {}Finally, we have \begin{equation} \zeta^{}_{\mathbb Q({\scriptscriptstyle \sqrt{2}})}(2) \; = \; \frac{\pi^4}{48 \sqrt{2}} \quad , \quad \zeta^{}_{\mathbb Q({\scriptscriptstyle \sqrt{2}})}(4) \; = \; \frac{11 \pi^8}{69120 \sqrt{2}} \, . \end{equation} \clearpage \vspace{5mm}
\section{Introduction} \renewcommand{\theequation}{1.\arabic{equation}} Sin-Gordon theory is one of the most studied examples of exactly integrable relativistic QFT in two dimensions. Its action \begin{equation} {\cal S}_{SG}=\int d^{2}x\left\{ \frac{1}{16\pi }\partial _{\nu }\varphi \partial ^{\nu }\varphi +2\mu \cos \beta \varphi \right\} \label{0.5} \end{equation} may be viewed as a perturbation of the free massless Bose field by the relevant operator $\cos \beta \varphi $ with scaling dimension $d=2\beta ^{2}<2.$ It is convenient to normalize exponential fields by the condition, that in UV limit their two point functions approach to those of $c=1$ conformal free Bose field theory \begin{equation} \left\langle e^{ia\varphi \left( x\right) }e^{-ia\varphi \left( y\right) }\right\rangle _{SG}\rightarrow \left| x-y\right| ^{-4a^{2}}\quad as\ \left| x-y\right| \rightarrow 0. \label{0.6} \end{equation} The on-shell properties of this theory i.e. the mass spectrum and the S-matrix are well known \cite{ZamZam1}. The lightest particles of the theory are solitons and antisolitons in term of which sin-Gordon theory has an equivalent Lagrangian formulation with the action of Massive Thirring Model (MTM) \cite{Col} \begin{equation} {\cal S}_{MTM}=\int d^{2}x\left\{ i\overline{\Psi }\gamma ^{\nu }\partial _{\nu }\Psi -M\ \overline{\Psi }\Psi -\frac{g}{2}\left( \overline{\Psi } \gamma ^{\nu }\Psi \right) \left( \overline{\Psi }\gamma _{\nu }\Psi \right) \right\} , \label{0.7} \end{equation} where $\overline{\Psi }$, $\Psi $ are two component Dirac spinors and the corresponding (anti-) particles are identified with the (anti-) solitons of ( \ref{0.5}). The famous Coleman relations \begin{equation} \frac{g}{\pi }=\frac{1}{2\beta ^{2}}-1;\ J^{\nu }\equiv \overline{\Psi } \gamma ^{\nu }\Psi =-\frac{\beta }{2\pi }\epsilon ^{\nu \mu }\partial _{\mu }\varphi \label{0.8} \end{equation} serve as a dictionary between bosonic and fermionic languages. More recently comparing the results of the thermodynamic Bethe-Ansatz analyses with those of the Conformal Perturbation Theory an exact relation between the soliton mass $M$ and the perturbation parameter $\mu $ is established \cite{AlZam} \begin{equation} \mu =\frac{\Gamma \left( \beta ^{2}\right) }{\pi \Gamma \left( 1-\beta ^{2}\right) }\left[ \frac{M\sqrt{\pi }\Gamma \left( \frac{1+\xi }{2}\right) }{2\Gamma \left( \frac{\xi }{2}\right) }\right] ^{2-2\beta ^{2}}, \label{0.9} \end{equation} where \begin{equation} \xi =\frac{\beta ^{2}}{1-\beta ^{2}}. \label{0.10} \end{equation} Starting from the expressions for the special cases $\beta \rightarrow 0$ (semiclassical limit) and $\beta ^{2}=1/2$ (free fermion case) S.Lukyanov and A.Zamolodchikov in \cite{LukZam} the following general formula for the Vacuum Expectation Value (VEV) $G_{a}=\left\langle \exp ia\varphi \left( 0\right) \right\rangle $ have conjectured \begin{eqnarray} &&G_{a}=\left( \frac{m\Gamma \left( \frac{1+\xi }{2}\right) \Gamma \left( 1-\frac{\xi }{2}\right) }{4\sqrt{\pi }}\right) ^{2a^{2}}\times \label{0.1} \\ &&\exp \left\{ \int\limits_{0}^{\infty }\frac{dt}{t}\left[ \frac{\sinh ^{2}\left( 2a\beta t\right) }{2\sinh \beta ^{2}t\sinh t\cosh \left( \left( 1-\beta ^{2}\right) t\right) }-2a^{2}e^{-2t}\right] \right\} . \nonumber \end{eqnarray} In the subsequent papers \cite{FatLukZamZam1}, \cite{FatLukZamZam2} some convincing arguments have been presented, showing that the VEV's $\widetilde{ G}_{a}=\left\langle \exp a\varphi \left( 0\right) \right\rangle _{sh-G}$ in sinh-Gordon theory, the action of which formally can be obtained simply replacing $\beta \rightarrow ib$ in (\ref{0.5}), obey the functional relations \begin{equation} \widetilde{G}_{a}=R\left( a\right) \widetilde{G}_{Q-a}=\widetilde{G}_{-a}, \label{0.2} \end{equation} where $Q=b+1/b$ and $R\left( a\right) $ is related to the Liouville reflection amplitude $S\left( p\right) $ \cite{ZamZam2} \begin{equation} R\left( \frac{Q}{2}+ip\right) =S\left( p\right) =-\left( \frac{\pi \mu \Gamma \left( b^{2}\right) }{\Gamma \left( 1-b^{2}\right) }\right) ^{-\frac{ 2ip}{b}}\frac{\Gamma \left( 1+2ip/b\right) \Gamma \left( 1+2ipb\right) }{ \Gamma \left( 1-2ip/b\right) \Gamma \left( 1-2ipb\right) }\ . \label{0.3} \end{equation} It is natural to expect that $\widetilde{G}_{a}$ can be obtained from $G_{a}$ making the analytic continuation \begin{equation} \beta \rightarrow ib;\ a\rightarrow -ia. \label{0.4} \end{equation} And indeed it can be shown that after substitution (\ref{0.4}), the expression (\ref{0.1}) obeys the relation (\ref{0.2}). Unfortunately the functional relations (\ref{0.2}) alone are not sufficient to find $ \widetilde{G}_{a}$ uniquely: the multiplication by any even, periodic with period $Q$ function of $a$ gives a different solution to (\ref{0.2}). However, $\widetilde{G}_{a}$ obtained from (\ref{0.1}) is the only {\it meromorphic }solution, obeying the extra requirement of "minimality" (i.e. the condition that only the poles and zeros, imposed by the functional relations (\ref{0.2}) are allowed). So, any independent test of the Lukyanov-Zamolodchikov formula (\ref{0.1}) will support the minimality assumption as well. This is important also because there are other interesting models for which some functional relations like (\ref{0.2}) are hold and the minimality condition makes it possible to find exact VEV's. In this article a perturbation theory based on the angular quantization \cite {Luk} of the MTM (\ref{0.7}) is developed, using which Lukyanov-Zamolodchikov formula (\ref{0.1}) near the free fermion point in first order over the MTM coupling constant $g$ is tested. The section 2 is devoted to the angular quantization of the MTM (\ref{0.7}). A particular attention is payed to the local field product regularization procedure, which has some additional features in comparison with the case of the ordinary quantization in Cartesian coordinates. In section 3 we calculate VEV $\left\langle \exp ia\varphi \left( 0\right) \right\rangle $ near the free fermion point. It appears that the Hankel-transform is a very useful tool to carry out this calculation. Some of the related mathematical details are presented in Appendix A. The choices we have made to regularize the traces over the fermionic Fock space and the local field product, result in a finite multiplicative renormalization of the field $\exp ia\varphi \left( 0\right) $. The corresponding renormalization factor is calculated using the methods of Boundary CFT \cite {Cardy} in Appendix B. The final expression we obtained for the expansion of the VEV $\left\langle \exp ia\varphi \left( 0\right) \right\rangle $ up to the first order in MTM coupling constant $g$ is in complete agreement with the Lukyanov-Zamolodchikov conjecture (\ref{0.1}). \vspace{0.5cm} \setcounter{equation}{0} \section{Angular Quantization of the Massive Thirring Model} \renewcommand{\theequation}{2.\arabic{equation}} It is convenient to use the ciral representation of the Dirac matrices \begin{equation} \gamma ^{0}=\sigma _{2}=\left( \begin{array}{ll} 0 & -i \\ i & 0 \end{array} \right) ,\ \gamma ^{1}=-i\sigma _{1}=\left( \begin{array}{ll} 0 & -i \\ -i & 0 \end{array} \right) \label{1.1} \end{equation} and denote the components of Dirac spinors as \begin{equation} \Psi \equiv \left( \begin{array}{l} \psi _{L} \\ \psi _{R} \end{array} \right) ,\ \overline{\Psi }\equiv \Psi ^{\dagger }\gamma ^{0}. \label{1.2} \end{equation} In this notations the action (\ref{0.7}) in Euclidean space reads \begin{eqnarray} {\cal A}_{MTM} &=&\int d^{2}z\ 2\left[ \psi _{R}^{\dagger }\partial \psi _{R}+\psi _{L}^{\dagger }\overline{\partial }\psi _{L}-\frac{iM}{2}\left( \psi _{L}^{\dagger }\psi _{R}-\psi _{R}^{\dagger }\psi _{L}\right) \right. \nonumber \\ &&\ \qquad \qquad \qquad \qquad \qquad \qquad \qquad \quad \left. +g\psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right] , \label{1.3} \end{eqnarray} where $z=\tau +ix^{1}$, $\overline{z}=\tau -ix^{1}$ are complex coordinates on the plane ($\tau =ix^{0}$ is the Euclidean time ), $\partial \equiv \partial /\partial z$, $\overline{\partial }\equiv \partial /\partial \overline{z}$ and the measure $d^{2}z\equiv dx^{1}d\tau $. As the VEV's of local fields $\left\langle e^{ia\varphi \left( 0\right) }\right\rangle $ have rotational symmetry, it is natural to use the conformal polar coordinates $\eta $, $\theta $ defined by \begin{equation} z\equiv e^{\eta +i\theta };\quad \overline{z}\equiv e^{\eta -i\theta } \label{1.4} \end{equation} and treat $\eta $, $\theta $ as space and (Euclidean) time respectively. Under the coordinate transformation (\ref{1.4}) the Fermi fields transform as follows: \begin{equation} \psi _{L}\left( z,\overline{z}\right) = e^{-\frac{i\pi }{4}-\frac{ \eta +i\theta }{2}}\psi _{L}\left( \eta ,\theta \right) ;\ \psi _{R}\left( z, \overline{z}\right) = e^{\frac{i\pi }{4}-\frac{\eta -i\theta }{2} }\psi _{R}\left( \eta ,\theta \right) , \label{1.5} \end{equation} and similarly for the conjugate fields $\psi _{L}^{\dagger }$, $\psi _{R}^{\dagger }$. In this coordinates the action (\ref{1.3}) takes the form: \begin{eqnarray} {\cal A}_{MTM} &=&\int\limits_{0}^{2\pi }d\theta \int\limits_{-\infty }^{\infty }d\eta \ \left[ \psi _{L}^{\dagger }\left( \partial _{\theta }-i\partial _{\eta }\right) \psi _{L}+\psi _{R}^{\dagger } \left( \partial _{\theta}+i\partial _{\eta }\right) \psi _{R}-\right. \nonumber \\ \quad \qquad \qquad &&\qquad \quad \left. iMe^{\eta }\left( \psi _{L}^{\dagger }\psi _{R}-\psi _{R}^{\dagger }\psi _{L}\right) +2g\psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right] . \label{1.6} \end{eqnarray} The usual canonical quantization yields the following ''equal time'' anti-commutation relations: \begin{equation} \left\{ \psi _{L}\left( \eta \right) ,\psi _{L}^{\dagger }\left( \eta ^{\prime }\right) \right\} =\delta \left( \eta -\eta ^{\prime }\right) ,\quad \left\{ \psi _{R}\left( \eta \right) ,\psi _{R}^{\dagger }\left( \eta ^{\prime }\right) \right\} =\delta \left( \eta -\eta ^{\prime }\right) . \label{1.7} \end{equation} From (\ref{1.6}) one deduces that the Hamiltonian defining the evolution along $\theta $ is given by \begin{eqnarray} \mbox{{\bf K}} &=&\int\limits_{-\infty }^{\infty }d\eta \left[ -\psi _{L}^{\dagger }i\partial _{\eta }\psi _{L}+\psi _{R}^{\dagger }i\partial _{\eta }\psi _{R}-iMe^{\eta }\left( \psi _{L}^{\dagger }\psi _{R}-\psi _{R}^{\dagger }\psi _{L}\right) +\right. \nonumber \\ &&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \quad \left. 2g\psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right] . \label{1.8} \end{eqnarray} To regularize the theory, following \cite{LukZam} let us restrict the range of the ''space'' coordinate $\eta $ to the semi-infinite box $\eta \in \left[ \log \varepsilon ;\infty \right) $ with $M\varepsilon \ll 1$ and impose the boundary conditions \begin{equation} \psi _{L}+\psi _{L}^{\dagger }\mid _{\eta =\log \varepsilon }=\psi _{R}+\psi _{R}^{\dagger }\mid _{\eta =\log \varepsilon }=0. \label{1.9} \end{equation} As usual, to develope perturbation theory in interaction picture, one first has to diagonalize the quadratic part of the Hamiltonian (\ref{1.8}). This can be achieved using the decomposition \cite{LukZam} \begin{eqnarray} &&\psi _{L}\left( \eta ,\theta \right) =\sum\limits_{\upsilon \in {\cal N} _{\varepsilon }}\frac{1}{\sqrt{2\pi \rho \left( \nu \right) }}\ c_{\nu }u_{\nu }\left( \eta \right) e^{-\nu \theta }\ , \nonumber \\ \ &&\psi _{R}\left( \eta ,\theta \right) =\sum\limits_{\upsilon \in {\cal N} _{\varepsilon }}\frac{1}{\sqrt{2\pi \rho \left( \nu \right) }}\ c_{\nu }\upsilon _{\nu }\left( \eta \right) e^{-\nu \theta }\ , \nonumber \\ &&\psi _{L}^{\dagger }\left( \eta ,\theta \right) =\sum\limits_{\upsilon \in {\cal N}_{\varepsilon }}\frac{1}{\sqrt{2\pi \rho \left( \nu \right) }}\ c_{\nu }^{\dagger }u_{\nu }^{*}\left( \eta \right) e^{\nu \theta }\ , \nonumber \\ &&\ \psi _{R}^{\dagger }\left( \eta ,\theta \right) =\sum\limits_{\upsilon \in {\cal N}_{\varepsilon }}\frac{1}{\sqrt{2\pi \rho \left( \nu \right) }}\ c_{\nu }^{\dagger }\upsilon _{\nu }^{*}\left( \eta \right) e^{\nu \theta }\ , \label{1.10} \end{eqnarray} where the set of admissable $\nu$'s ${\cal N}_{\varepsilon }$ and the density of states $\rho \left( \nu \right) $ are specified below (see (\ref{1.14}), (\ref{1.14'})) and the wave functions \cite{LukZam} \begin{equation} \left( \begin{array}{l} u_{\nu }\left( \eta \right) \\ v_{\nu }\left( \eta \right) \end{array} \right) =\frac{\sqrt{2M}e^{\frac{\eta }{2}}}{\Gamma \left( \frac{1}{2}-i\nu \right) }\left( \frac{M}{2}\right) ^{-i\nu }\left( \begin{array}{l} K_{\frac{1}{2}-i\nu }\left( Me^{\eta }\right) \\ K_{\frac{1}{2}+i\nu }\left( Me^{\eta }\right) \end{array} \right) , \label{1.11} \end{equation} are solutions of the free Dirac equation ($K_{\nu }\left( x\right) $ is the MacDonald function). The wave functions (\ref{1.11}) have the asymptotic behavior \begin{equation} \left( \begin{array}{l} u_{\nu }\left( \eta \right) \\ v_{\nu }\left( \eta \right) \end{array} \right) \rightarrow \left( \begin{array}{l} 1 \\ 0 \end{array} \right) e^{i\nu \eta }+S_{F}\left( \nu \right) \left( \begin{array}{l} 0 \\ 1 \end{array} \right) e^{-i\nu \eta }\ \ as\ \eta \rightarrow -\infty \ , \label{1.12} \end{equation} where \begin{equation} S_{F}\left( \nu \right) =\left( \frac{M}{2}\right) ^{-2i\nu }\frac{\Gamma \left( \frac{1}{2}+i\nu \right) }{\Gamma \left( \frac{1}{2}-i\nu \right) } \label{1.13} \end{equation} is the fermion scattering amplitude off the ''mass barrier'' \cite{LukZam}. It follows from (\ref{1.9}), (\ref{1.12}) that the set ${\cal N} _{\varepsilon }$ of admissable $\nu $'s over which the sum in (\ref{1.10}) is carried out consists of the solutions of the equation \begin{equation} 2\pi \left( n+\frac{1}{2}\right) =2\nu \log \frac{1}{\varepsilon }+\frac{1}{ i }\log S_{F}\left( \nu \right) , \label{1.14} \end{equation} where $n$ is arbitrary integer. Therefore the density of states is given by \begin{equation} \rho \left( \nu \right) =\frac{dn}{d\nu }=\frac{1}{\pi }\log \frac{1}{ \varepsilon }+\frac{1}{2\pi i}\log ^{^{\prime }}S_{F}\left( \nu \right) , \label{1.14'} \end{equation} Below we'll use the notation ${\cal N}_{\varepsilon }^{+}\subset {\cal N} _{\varepsilon }$ for the subset of positive $\nu $'s. The operators $c_{\nu } $, $c_{\nu }^{\dagger }$ satisfy the anti-commutation relations \begin{eqnarray} \left\{ c_{\nu },c_{\nu ^{\prime }}\right\} &=&\left\{ c_{\nu }^{\dagger },c_{\nu ^{\prime }}^{\dagger }\right\} =0, \nonumber \\ \left\{ c_{\nu },c_{\nu ^{\prime }}^{\dagger }\right\} &=&\delta _{\nu ,\nu ^{\prime }}. \label{1.15} \end{eqnarray} In terms of this operators the free part of the Hamiltonian (\ref{1.8}) takes a very simple form \begin{equation} {\bf K}_{0}=\sum\limits_{\nu \in {\cal N}_{\varepsilon }^{+}}\nu \left( c_{\nu }^{\dagger }c_{\nu }+c_{-\nu }c_{-\nu }^{\dagger }\right) , \label{1.16} \end{equation} which shows that $c_{\nu }^{\dagger }$, $c_{-\nu }$ ($c_{\nu },$ $c_{-\nu }^{\dagger }$) are fermion and anti-fermion creation (annihilation) operators. Therefore the vacuum state $\left| 0\right\rangle $ can be defined by \begin{equation} c_{\nu }\left| 0\right\rangle =c_{-\nu }^{\dagger }\left| 0\right\rangle =0,\quad \nu \in {\cal N}_{\varepsilon }^{+}, \label{1.17} \end{equation} and the Hilbert space of states ${\cal H}$ is spanned over the base vectors \begin{equation} \prod\limits_{\nu \in {\cal N}_{\varepsilon }^{+}}c_{\nu }^{\dagger n_{\nu }}c_{-\nu }^{\overline{n}_{\nu }}\left| 0\right\rangle , \label{1.18} \end{equation} where $n_{\nu }\in \left\{ 0,1\right\} $ ($\overline{n}_{\nu }\in \left\{ 0,1\right\} $) are the occupation numbers of fermions (anti-fermions) with energy $\nu $. Standard arguments show, that the expectation value of any quantity $ \left\langle {\bf X}\right\rangle $ in interacting theory (\ref{1.8}) may be evaluated using the formula \begin{eqnarray} \left\langle {\bf X}\right\rangle \equiv &&\frac{\int D\Psi D\overline{\Psi } e^{-{\cal A}_{MTM}}{\bf X}}{\int D\Psi D\overline{\Psi }e^{-{\cal A}_{MTM}}}= \nonumber \\ &&\frac{{\it Tr}_{{\cal H}}\left[ e^{-2\pi {\bf K}_{0}}T\left( e^{-\int {\bf K}_{I}\left( \eta ,\theta \right) d\eta d\theta }{\bf X}\right) \right] }{ {\it Tr}_{{\cal H}}\left[ e^{-2\pi {\bf K}_{0}}T\left( e^{-\int {\bf K} _{I}\left( \eta ,\theta \right) d\eta d\theta }\right) \right] }\ , \label{1.19} \end{eqnarray} where ${\bf X}$ in second line of (\ref{1.19}) as well as the interaction term of the Hamiltonian \begin{equation} {\bf K}_{I}=2g \int\limits_{\log \varepsilon }^{\infty }\ N\left( \psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right) d\eta \label{1.20} \end{equation} is taken in interaction picture and $T$ indicates the time ordering operation. In (\ref{1.20}) we introduced the notation $N\left( \cdots \right) $ for the suitably regularized product of the local operators in coinciding points (see below). Appearance of trace instead of conventional vacuum matrix element in (\ref{1.19}) is due to compactification of the ''time'' $\theta $. Now let us turn to the regularization procedure of the products of local operators at the coinciding points. In the ordinary case of non-compactified time one simply implies normal ordering prescription which is well known to be equivalent to the suppression of all contractions among the fields inside normal ordering symbol. In contrary to the vacuum matrix element, the trace of normal ordered product of creation and annihilation operators doesn't vanish, therefore the analogues regularization in the case of compactified time is slightly changed. It is easy to see that in this case the correctly regularized perturbing operator $N\left( \psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right) $ is given by \begin{eqnarray} &&N\left( \psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right) =\psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}-\left\langle \psi _{L}^{\dagger }\psi _{L}\right\rangle _{0}\psi _{R}^{\dagger }\psi _{R}-\left\langle \psi _{R}^{\dagger }\psi _{R}\right\rangle _{0}\psi _{L}^{\dagger }\psi _{L}+ \nonumber \\ &&\qquad \qquad \qquad \qquad \quad \left\langle \psi _{L}^{\dagger }\psi _{R}\right\rangle _{0}\psi _{R}^{\dagger }\psi _{L}+\left\langle \psi _{R}^{\dagger }\psi _{L}\right\rangle _{0}\psi _{L}^{\dagger }\psi _{R}+ \nonumber \\ &&\qquad \qquad \qquad \qquad \quad \left\langle \psi _{L}^{\dagger }\psi _{L}\right\rangle _{0}\left\langle \psi _{R}^{\dagger }\psi _{R}\right\rangle _{0}-\left\langle \psi _{L}^{\dagger }\psi _{R}\right\rangle _{0}\left\langle \psi _{R}^{\dagger }\psi _{L}\right\rangle _{0}= \nonumber \\ &&:\psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}:+\left\langle :\psi _{L}^{\dagger }\psi _{R}:\right\rangle _{0}:\psi _{R}^{\dagger }\psi _{L}:+\left\langle :\psi _{R}^{\dagger }\psi _{L}:\right\rangle _{0}:\psi _{L}^{\dagger }\psi _{R}:- \nonumber \\ &&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \left\langle :\psi _{L}^{\dagger }\psi _{R}:\right\rangle _{0}\left\langle :\psi _{R}^{\dagger }\psi _{L}:\right\rangle _{0}, \label{1.21} \end{eqnarray} where \begin{equation} \left\langle {\bf X}\right\rangle _{0}=\frac{{\it Tr}_{{\cal H}}e^{-2\pi {\bf K}_{0}}{\bf X}}{{\it Tr}_{{\cal H}}e^{-2\pi {\bf K}_{0}}}\ , \label{1.22} \end{equation} for any operator ${\bf X}$, $::$ denotes the ordinary normal ordering with respect to the mode decomposition (\ref{1.10}) and in the second equality of (\ref{1.21} ) we have taken into account that \begin{equation} \left\langle :\psi _{L}\psi _{R}:\right\rangle _{0}=\left\langle :\psi _{L}^{\dagger }\psi _{R}^{\dagger }:\right\rangle _{0}=\left\langle :\psi _{L}^{\dagger }\psi _{L}:\right\rangle _{0}=\left\langle :\psi _{R}^{\dagger }\psi _{R}:\right\rangle _{0}=0. \label{1.24} \end{equation} Note that, while to give a proper mining to the first expression for the perturbing operator $N\left( \psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right) $ (see (\ref{1.21}) ) one has to apply point splitting, the separate terms of the second expression are already finite. \vspace{0.5cm} \setcounter{equation}{0} \section{VEV's of Exponential Fields} \renewcommand{\theequation}{3.\arabic{equation}} The Dirac fields $\Psi \left( z,\overline{z}\right) $, $\overline{\Psi } \left( z,\overline{z}\right) $ have non-trivial monodromy with respect to the exponential fields $\exp ia\varphi \left( 0\right) $ \begin{eqnarray} \Psi \left( z,\overline{z}\right) *e^{ia\varphi \left( 0\right) } &=&e^{i \frac{2\pi a}{\beta }}\Psi \left( z,\overline{z}\right) e^{ia\varphi \left( 0\right) }\ , \nonumber \\ \overline{\Psi }\left( z,\overline{z}\right) *e^{ia\varphi \left( 0\right) } &=&e^{-i\frac{2\pi a}{\beta }}\overline{\Psi }\left( z,\overline{z}\right) e^{ia\varphi \left( 0\right) }\ , \label{2.1} \end{eqnarray} where $*$ denotes the analytic continuation around the point $0$. The VEV $ \left\langle \exp ia\varphi \left( 0\right) \right\rangle $ can be expressed in terms of Grassmanian functional integral \cite{LukZam} \begin{equation} I\left( a\right) =\frac{\int\limits_{{\cal F}_{a}}\left[ {\cal D}\Psi {\cal D }\overline{\Psi }\right] e^{-{\cal A}_{MTM}}}{\int\limits_{{\cal F} _{0}}\left[ {\cal D}\Psi {\cal D}\overline{\Psi }\right] e^{-{\cal A}_{MTM}}} , \label{2.2} \end{equation} where the functional integration in the numerator is carried out over the space ${\cal F}_{a}$ of twisted field configurations with monodromy (\ref {2.1}). In angular quantization picture the insertion of the operator $\exp ia\varphi \left( 0\right) $ changes the boundary conditions along ''time'' direction so that the Hilbert space remains untouched but the Hamiltonian due to (\ref{2.1}) acquires an additional term $ -ia{\bf Q}/\beta $, where \begin{equation} {\bf Q}=\sum\limits_{\nu \in {\cal N}_{\varepsilon }^{+}}\left( c_{\nu }^{\dagger }c_{\nu }-c_{-\nu }c_{-\nu }^{\dagger }\right) \label{2.3} \end{equation} is the fermion charge operator. Thus, due to (\ref{1.19}), for the regularized version of the functional integral (\ref{2.2}) we have \begin{equation} I_{\varepsilon }\left( a,g\right) =\frac{{\it Tr}_{{\cal H}}\left[ e^{-2\pi {\bf K}_{0}+\frac{2\pi ia}{\beta }{\bf Q}}T\left( e^{-\int {\bf K}_{I}\left( \eta ,\theta \right) d\eta d\theta }\right) \right] }{{\it Tr}_{{\cal H} }\left[ e^{-2\pi {\bf K}_{0}}T\left( e^{-\int {\bf K}_{I}\left( \eta ,\theta \right) d\eta d\theta }\right) \right] }{\it \ }. \label{2.4} \end{equation} The VEV $\left\langle \exp ia\varphi \left( 0\right) \right\rangle $ can be expressed in terms of $I_{\varepsilon }\left( a,g\right) $ as \begin{equation} \left\langle e^{ia\varphi \left( 0\right) }\right\rangle =\lim\limits_{\varepsilon \rightarrow 0}Z^{-1}\varepsilon ^{-2a^{2}}I_{\varepsilon }\left( a,g\right) , \label{2.5} \end{equation} where $Z$ is some renormalization factor. This point, as well as the appearance of the factor $\varepsilon ^{-2a^{2}}$, which has purely CFT origin, will be discussed later on. The main goal of this paper is the evaluation of (\ref{2.4}) and (\ref{2.5}) perturbatively up to the linear over $g$ terms \begin{equation} I_{\varepsilon }\left( a,g\right) =I_{\varepsilon }\left( a,0\right) \left( 1+gI_{\varepsilon }^{1}\left( a\right) +{\cal O}\left( g^{2}\right) \right) . \label{2.6} \end{equation} The calculation of \begin{equation} I_{\varepsilon }\left( a,0\right) =\frac{{\it Tr}_{{\cal H}}\left[ e^{-2\pi {\bf K}_{0}+2\pi ia\sqrt{2}{\bf Q}}\right] }{{\it Tr}_{{\cal H}}\left[ e^{-2\pi {\bf K}_{0}}\right] } \label{2.7} \end{equation} is carried out in \cite{LukZam} and the result is \begin{eqnarray} &&I_{\varepsilon }\left( a,0\right) =\varepsilon ^{2a^{2}}\left\langle e^{ia\varphi \left( 0\right) }\right\rangle \mid _{g=0}= \nonumber \\ &&\left( \frac{M\varepsilon }{2}\right) ^{2a^{2}}\exp \left\{ \int\limits_{0}^{\infty }\frac{dt}{t}\left[ \frac{\sinh ^{2}\left( \sqrt{2} at\right) }{\sinh ^{2} t}-2a^{2}e^{-2t}\right] \right\} , \label{2.22} \end{eqnarray} so that we'll concentrate our attention on the second term \begin{eqnarray} I_{\varepsilon }^{1}\left( a\right) &=&\frac{a}{2\pi }\partial _{a}\log I_{\varepsilon }\left( a,0\right) - \nonumber \\ &&\frac{4\pi {\it Tr}_{{\cal H}}\left[ e^{-2\pi {\bf K}_{0}+2\pi ia\sqrt{2} {\bf Q}}\int\limits_{\log \varepsilon }^{\infty }d\eta N\left( \psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right) \right] }{{\it Tr}_{{\cal H}}\left[ e^{-2\pi {\bf K}_{0}+2\pi ia\sqrt{2}{\bf Q}}\right] }. \label{2.8} \end{eqnarray} Using mode decomposition (\ref{1.10}) and evaluating traces over ${\cal H}$ in the bases (\ref{1.18}) for (\ref{2.8}) we obtain \begin{eqnarray} &&I_{\varepsilon }^{1}\left( a\right) =\frac{a}{2\pi }\partial _{a}\log I_{\varepsilon }\left( a,0\right) +\sum\limits_{\nu _{1},\nu _{2}\in {\cal N} _{\varepsilon }^{+}}\left\{ \frac{\cosh \pi \nu _{1}\cosh \pi \nu _{2}}{\pi ^{4}\rho \left( \nu _{1}\right) \rho \left( \nu _{2}\right) }\times \right. \nonumber \\ &&\int\limits_{M\varepsilon }^{\infty }\left[ A_{\nu _{1}}A_{\nu _{2}}\left( \left| K_{\frac{1}{2}+i\nu _{1}}\left( x\right) \right| ^{2}\left| K_{\frac{1 }{2}+i\nu _{2}}\left( x\right) \right| ^{2}-K_{\frac{1}{2}+i\nu _{1}}^{2}\left( x\right) K_{\frac{1}{2}-i\nu _{2}}^{2}\left( x\right) \right) +\right. \nonumber \\ &&\qquad A_{\nu _{1}}A_{\nu _{2}}^{*}\left( K_{\frac{1}{2}+i\nu _{1}}^{2}\left( x\right) K_{\frac{1}{2}+i\nu _{2}}^{2}\left( x\right) -\left| K_{\frac{1}{2}+i\nu _{1}}\left( x\right) \right| ^{2}\left| K_{\frac{ 1}{2}+i\nu _{2}}\left( x\right) \right| ^{2}\right) - \nonumber \\ &&\qquad \qquad \qquad \qquad A_{\nu _{1}}^{0}\left( A_{\nu _{2}}-\frac{1}{2} A_{\nu _{2}}^{0}\right) \left( K_{\frac{1}{2}+i\nu _{1}}^{2}\left( x\right) -K_{\frac{1}{2}-i\nu _{1}}^{2}\left( x\right) \right) \times \nonumber \\ &&\qquad \qquad \qquad \qquad \left. \left. \left( K_{\frac{1}{2}+i\nu _{2}}^{2}\left( x\right) -K_{\frac{1}{2}-i\nu _{2}}^{2}\left( x\right) \right) +{\bf c.c.}\right] xdx\right\} \ , \label{2.9} \end{eqnarray} where \begin{equation} A_{\nu }=\frac{e^{-2\pi \nu +2i\sqrt{2}\pi a}}{1+e^{-2\pi \nu +2i\sqrt{2}\pi a}}, \label{2.10} \end{equation} and $A_{\nu }^{0}\equiv A_{\nu }\mid _{a=0}$. The following formulae for the integrals over $x$ included in (\ref{2.9}) are proved in Appendix A (below and later on we'll omit vanishing in the limit $\varepsilon \rightarrow 0$ terms) \begin{eqnarray} &&\frac{\cosh \pi \nu _{1}\cosh \pi \nu _{2}}{\pi ^{2}}\int\limits_{M \varepsilon }^{\infty }\left| K_{\frac{1}{2}+i\nu _{1}}\left( x\right) \right| ^{2}\left| K_{\frac{1}{2}+i\nu _{2}}\left( x\right) \right| ^{2}xdx= \nonumber \\ &&\int\limits_{0}^{\infty }\left[ \frac{\sin ^{2}\nu _{1}t+\sin ^{2}\nu _{2}t }{2\sinh t}+\frac{\cos 2\nu _{1}t\cos 2\nu _{2}t-1}{2\sinh 2t}\right] dt- \nonumber \\ &&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \frac{\gamma +2\log 2+\log M\varepsilon }{4}\ , \label{2.11} \end{eqnarray} \begin{eqnarray} &&\frac{\cosh \pi \nu _{1}\cosh \pi \nu _{2}}{\pi ^{2}}\int\limits_{M \varepsilon }^{\infty }K_{\frac{1}{2}+i\nu _{1}}^{2}\left( x\right) K_{\frac{ 1}{2}+i\nu _{2}}^{2}\left( x\right) xdx= \nonumber \\ &&\int\limits_{0}^{\infty }\left[ \frac{\sinh \left( 1+2i\nu _{1}\right) t\sinh \left( 1+2i\nu _{2}\right) t}{2\sinh 2t}-\frac{1}{4}e^{2i\left( \nu _{1}+\nu _{2}\right) t}\right] dt+ \nonumber \\ &&\qquad \qquad \frac{\gamma \left( \frac{1}{2}+i\nu _{1}\right) \gamma \left( \frac{1}{2}+i\nu _{2}\right)\left( \frac{2}{M\varepsilon }\right) ^{2i\left( \nu _{1}+\nu _{2}\right) } -1}{8i\left( \nu _{1}+\nu _{2}\right) } \ , \label{2.12} \end{eqnarray} where $\gamma =0.577216\cdots $ is the Euler constant and \begin{equation} \gamma \left( x\right) \equiv \frac{\Gamma \left( x\right) }{\Gamma \left( 1-x\right) }\ . \label{2.25} \end{equation} Let us imagine that $x$ integration in (\ref{2.9}) with the help of (\ref {2.11}) and (\ref{2.12}) is already performed. Then the resulting expression can be represented as a sum of two parts \begin{equation} I_{\varepsilon }^{1}\left( a\right) =\sum\limits_{\nu _{1},\nu _{2}\in {\cal N}_{\varepsilon }^{+}}\frac{1}{\rho \left( \nu _{1}\right) \rho \left( \nu _{2}\right) }\left[ {\LARGE i}\right] +\sum\limits_{\nu _{1},\nu _{2}\in {\cal N}_{\varepsilon }^{+}}\frac{1}{\rho \left( \nu _{1}\right) \rho \left( \nu _{2}\right) }\left[ {\Large ii}\right] , \label{2.13} \end{equation} where in the second part, symbolically denoted as $\left[ ii\right] $, are collected all the terms induced by the term \begin{equation} \frac{\gamma \left( \frac{1}{2}+i\nu _{1}\right) \gamma \left( \frac{1}{2} +i\nu _{2}\right)\left( \frac{2}{M\varepsilon }\right) ^{2i\left( \nu _{1}+ \nu _{2}\right) } -1}{8i\left( \nu _{1}+\nu _{2}\right) } \label{2.23} \end{equation} of the equation (\ref{2.12}). As the $\nu _{1}$, $\nu _{2}$ dependence of terms collected in the remaining part $\left[ i\right] $ (in contrary to those of $\left[ ii\right] $ ) are free of rapid , comparable with $\log 1/M\varepsilon $ frequency oscillations, it is safe to make replacement \begin{equation} \sum\limits_{\nu _{1},\nu _{2}\in {\cal N}_{\varepsilon }^{+}}\frac{1}{\rho \left( \nu _{1}\right) \rho \left( \nu _{2}\right) }\left[ {\Large i}\right] \rightarrow \int\limits_{0}^{\infty }\int\limits_{0}^{\infty }d\nu _{1}d\nu _{2}\left[ {\Large i}\right] \ . \label{2.14} \end{equation} Surprisingly enough it is possible to factorize these integrals in such a way, that they can be performed explicitly using the formula (for proof see Appendix A) \begin{equation} Im\int\limits_{0}^{\infty }d\nu \frac{e^{-2\pi \nu +i\pi \alpha +2i\nu t}}{ 1+e^{-2\pi \nu +i\pi \alpha }}=\frac{1}{4t}-\frac{e^{-\alpha t}}{4\sinh t}\ . \label{2.26} \end{equation} The sum over $\nu _{1}$, $\nu _{2}$ in second term of (\ref{2.13}) also can be converted into the integrals provided one notices that (\ref{2.23}) is exactly zero when $\nu _{1},\nu _{2}\in {\cal N}_{\varepsilon }$ , $\nu _{1}+\nu _{2}\neq 0$ and is equal to $\pi \rho \left( \nu _{1}\right) /4$ (up to nonessential, finite in the $\varepsilon \rightarrow 0$ limit term) if $\nu _{1}+\nu _{2}=0$, so that (\ref{2.23}) can be effectively replaced by $\pi \delta \left( \nu _{1}+\nu _{2}\right) /4$. This leaves us with elementary one dimensional integrals. As a result of above described calculation for (\ref{2.9}) one obtains \begin{eqnarray} &&2\pi I_{\varepsilon }^{1}\left( a\right) =\pi \alpha \cot \frac{\pi \alpha }{2}-2+\frac{\alpha ^{2}}{2}\left( \psi \left( \frac{1}{2}\right) -\log 2\right) + \nonumber \\ &&\int\limits_{0}^{\infty }dt\left[ \frac{\sinh ^{2}\alpha t}{\sinh 2t\sinh ^{2}t}+\frac{\alpha ^{2}\left( 2\cosh t-1\right) }{\sinh 2t}-\frac{\alpha \sinh \alpha t}{2\sinh ^{2}t}+\right. \label{2.15} \\ &&\left. \frac{2\cosh t\sinh ^{4}\frac{\alpha }{2}t}{\sinh ^{3}t}+\frac{ \cosh \alpha t-1}{\sinh ^{2}t}-\frac{\sinh ^{2}\alpha t}{\sinh 2t}- \frac{\alpha ^{2} e^{-2t}}{2t}\right] \ , \nonumber \end{eqnarray} where $\psi \left( x\right) $ is the logarithmic derivative of the $\Gamma $ -function. In (\ref{2.15}) and later on we set \begin{equation} \alpha =2\sqrt{2}a\ . \label{2.24} \end{equation} Using the table of integrals presented at the end of Appendix A it is not difficult to perform $t$ integration too \begin{eqnarray} &&2\pi I_{\varepsilon }^{1}\left( a\right) =-\frac{1}{2}\psi \left( \frac{1}{ 2}\right) -1+\frac{\pi \alpha }{2}\cot \frac{\pi \alpha }{2}+\alpha ^{2}\left( \frac{1}{4}-\log 2\right) + \nonumber \\ &&\frac{\alpha ^{2}}{4}\left( \psi \left( \frac{\alpha }{2}\right) +\psi \left( -\frac{\alpha }{2}\right) \right) +\frac{1-\alpha ^{2}}{4}\left( \psi \left( \frac{1+\alpha }{2}\right) +\psi \left( \frac{1-\alpha }{2}\right) \right) . \label{2.16} \end{eqnarray} Now let us turn to the computation of the renormalization factor $Z$ in ( \ref{2.5}). If $\varepsilon $ is small enough we can split the region $ \left| z\right| \geq \varepsilon $ into two pieces by the circle $\left| z\right| =\widetilde{\varepsilon }$ with some $\widetilde{\varepsilon }$ satisfying the conditions \begin{equation} \log \frac{\widetilde{\varepsilon }}{\varepsilon }\gg 1;\quad \log M \widetilde{\varepsilon }\ll 1\, \label{2.17} \end{equation} so that inside the first region ${\it U}_{1}=\left\{ z;\varepsilon \leq \left| z\right| \leq \widetilde{\varepsilon }\right\} $ the theory is nearly conformal invariant and at the same time in the region ${\it U}_{2}=\left\{ z;\left| z\right| >\widetilde{\varepsilon }\right\} $ the influence of the boundary at $\left| z\right| =\varepsilon $ could be neglected. Note that in contrary to the region ${\it U}_{2}$ where the regularization prescription ( \ref{1.21}) is standard and in Cartesian coordinates transforms to the usual normal ordering , due to the influence of the boundary in region ${\it U} _{1} $ the interaction term of the Hamiltonian (\ref{1.20}) results in an extra multiplicative renormalization of the field $\exp ia\varphi $ (besides the usual charge renormalization $a_{r}=\left( 1-g/2\pi \right) a$ ). The actual computation of the renormalization constant $Z$ is significantly simplified owing to the existence of conformal invariance inside the region $ {\it U}_{1} $. As usual in CFT it is convenient to use radial quantization \cite{BPZ}. let us denote by $\left| B,a\right\rangle $ the boundary state \cite{Cardy} corresponding to the boundary conditions (\ref{1.9}) and belonging to the conformal family \cite{BPZ} of the state $\left| a\right\rangle =$ $\exp ia\varphi \left( 0\right) \left| 0\right\rangle $. Note that during the evaluation from $\varepsilon $ to $\widetilde{ \varepsilon }$, the state $\varepsilon ^{-2a^{2}}\left| B,a\right\rangle $ approaches to $\widetilde{\varepsilon }^{-2a^{2}}\left| a\right\rangle $ thus correctly imitating the insertion of the field $\exp ia\varphi \left( 0\right) $. This consideration makes transparent the appearance of the factor $\varepsilon ^{-2a^{2}}$ in (\ref{2.5}). It is not difficult to see that the renormalization factor $Z$ is given by \begin{equation} Z-1=-4\pi g\int\limits_{\varepsilon }^{\widetilde{\varepsilon }}\left\langle a\right| N\left( \psi _{L}^{\dagger }\psi _{L}\psi _{R}^{\dagger }\psi _{R}\right) \varepsilon ^{L_{0}+\overline{L}_{0}}\left| B,a\right\rangle \left| z\right| d\left| z\right| -\frac{g\alpha ^{2}}{4\pi }\log \frac{ \widetilde{\varepsilon }}{\varepsilon }\ . \label{2.18} \end{equation} Here and in what follows we take the Fermi fields in initial coordinates $z$ , $\overline{z}$ (i.e. the transformation (\ref{1.5}) is not applied). The operator $\varepsilon ^{L_{0}+\overline{L}_{0}}$ ($L_{0}$, $\overline{L}_{0}$ are the Virasoro generators) is included in (\ref{2.18}) to take into account that the boundary state $\left| B,a\right\rangle $ is associated to the circle $\left| z\right| =\varepsilon $. The second term in (\ref{2.18}) subtracts the contribution of the charge renormalization. In Appendix B we have presented the details of the computation of the matrix element included in (\ref{2.18}). Inserting (\ref{B6}) into (\ref{2.18}) and performing integration with the help of (\ref{A22}) we obtain \begin{equation} Z-1=-\frac{g}{2\pi }\left( 1-\frac{\pi \alpha }{2}\cot \frac{\pi \alpha }{2} \right) . \label{2.19} \end{equation} As it should be expected the choice of $\widetilde{\varepsilon }$ satisfying the conditions (\ref{2.17}) has no effect on the value of $Z$. Now taking into account (\ref{2.5}), (\ref{2.22}), (\ref{2.16}) and (\ref{2.19}) we can write down a final expression for the expansion of VEV up to linear over $g$ terms \begin{eqnarray} &&\frac{\left\langle e^{ia\varphi \left( 0\right) }\right\rangle }{\left\langle e^{ia\varphi \left( 0\right) }\right\rangle \mid _{g=0}}=1+\frac{g}{8\pi } \left[ -2\psi \left( \frac{1}{2}\right) +\alpha ^{2}\left( 1-4\log 2\right) +\right. \nonumber \\ &&\left. \alpha ^{2}\left( \psi \left( \frac{\alpha }{2}\right) +\psi \left( -\frac{\alpha }{2}\right) \right) +\left( 1-\alpha ^{2}\right) \left( \psi \left( \frac{1+\alpha }{2}\right) +\psi \left( \frac{1-\alpha }{2}\right) \right) \right] \nonumber \\ &&\qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \quad +O\left( g^{2}\right) . \label{2.21} \end{eqnarray} It is not difficult to check that (\ref{2.21}) exactly coincides with the expression which one obtains directly expanding Lukyanov-Zamolodchikov formula \cite{LukZam}. \vspace{0.5cm}
\section{Introduction} Networks of social interactions between individuals, groups, or organizations have some unusual topological properties which set them apart from most of the networks with which physics deals. They appear to display simultaneously properties typical both of regular lattices and of random graphs. For instance, social networks have well-defined locales in the sense that if individual~A knows individual~B and individual~B knows individual~C, then it is likely that A also knows C---much more likely than if we were to pick two individuals at random from the population and ask whether they are acquainted. In this respect social networks are similar to regular lattices, which also have well-defined locales, but very different from random graphs, in which the probability of connection is the same for any pair of vertices on the graph. On the other hand, it is widely believed that one can get from almost any member of a social network to any other via only a small number of intermediate acquaintances, the exact number typically scaling as the logarithm of the total number of individuals comprising the network. Within the population of the world, for example, it has been suggested that there are only about ``six degrees of separation'' between any human being and any other\cite{Milgram67}. This behavior is not seen in regular lattices but is a well-known property of random graphs, where the average shortest path between two randomly-chosen vertices scales as $\log N/\log z$, where $N$ is the total number of vertices in the graph and $z$ is the average coordination number\cite{Bollobas85}. Recently, Watts and Strogatz\cite{WS98} have proposed a model which attempts to mimic the properties of social networks. This ``small-world'' model consists of a network of vertices whose topology is that of a regular lattice, with the addition of a low density $\phi$ of connections between randomly-chosen pairs of vertices\cite{note1}. Watts and Strogatz showed that graphs of this type can indeed possess well-defined locales in the sense described above while at the same time possessing average vertex--vertex distances which are comparable with those found on true random graphs, even for quite small values of~$\phi$. In this paper we study in detail the behavior of the small-world model, concentrating particularly on its scaling properties. The outline of the paper is as follows. In Section~\ref{model} we define the model. In Section~\ref{length} we study the typical length-scales present in the model and argue that the model undergoes a continuous phase transition as the density of random connections tends to zero. We also examine the cross-over been large- and small-world behavior in the model, and the structure of ``neighborhoods'' of adjacent vertices. In Section~\ref{scaling} we derive a scaling form for the average vertex--vertex distance on a small-world graph and demonstrate numerically that this form is followed over a wide range of the parameters of the model. In Section~\ref{seceffdim} we calculate the effective dimension of small-world graphs and show that this dimension depends on the length-scale on which we examine the graph. In Section~\ref{percolation} we consider the properties of site percolation on these systems, as a model of the spread of information or disease through social networks. Finally, in Section~\ref{concs} we give our conclusions. \section{The small-world model} \label{model} The original small-world model of Watts and Strogatz, in its simplest incarnation, is defined as follows. We take a one-dimensional lattice of $L$ vertices with connections or bonds between nearest neighbors and periodic boundary conditions (the lattice is a ring). Then we go through each of the bonds in turn and independently with some probability $\phi$ ``rewire'' it. Rewiring in this context means shifting one end of the bond to a new vertex chosen uniformly at random from the whole lattice, with the exception that no two vertices can have more than one bond running between them, and no vertex can be connected by a bond to itself. In this model the average coordination number $z$ remains constant ($z=2$) during the rewiring process, but the coordination number of any particular vertex may change. The total number of rewired bonds, which we will refer to as ``shortcuts'', is $\phi L$ on average. For the purposes of analytic treatment the Watts--Strogatz model has a number of problems. One problem is that the distribution of shortcuts is not completely uniform; not all choices of the positions of the rewired bonds are equally probable. For example, configurations with more than one bond between a particular pair of vertices are explicitly forbidden. This non-uniformity of the distribution makes an average over different realizations of the randomness hard to perform. \begin{figure} \begin{center} \psfig{figure=onedim.ps,width=\columnwidth} \end{center} \caption{(a) An example of a small-world graph with $L=24$, $k=1$ and, in this case, four shortcuts. (b)~An example with $k=3$.} \label{onedim} \end{figure} A more serious problem is that one of the crucial quantities of interest in the model, the average distance between pairs of vertices on the graph, is poorly defined. The reason is that there is a finite probability of a portion of the lattice becoming detached from the rest in this model. Formally, we can represent this by saying that the distance from such a portion to a vertex elsewhere on the lattice is infinite. However, this means that the average vertex--vertex distance on the lattice is then itself infinite, and hence that the vertex--vertex distance averaged over all realizations is also infinite. For numerical studies such as those of Watts and Strogatz this does not present any substantial difficulties, but for analytic work it results in a number of quantities and expressions being poorly defined. Both of these problems can be circumvented by a slight modification of the model. In our version of the small-world model we again start with a regular one-dimensional lattice, but now instead of rewiring each bond with probability $\phi$, we add shortcuts between pairs of vertices chosen uniformly at random but we do not remove any bonds from the regular lattice. We also explicitly allow there to be more than one bond between any two vertices, or a bond which connects a vertex to itself. In order to preserve compatibility with the results of Watts and Strogatz and others, we add with probability $\phi$ one shortcut for each bond on the original lattice, so that there are again $\phi L$ shortcuts on average. The average coordination number is $z=2(1+\phi)$. This model is equivalent to the Watts--Strogatz model for small $\phi$, whilst being better behaved when $\phi$ becomes comparable to~1. Fig.~\ref{onedim}(a) shows one realization of our model for $L=24$. Real social networks usually have average coordination numbers $z$ significantly higher than~$2$, and we can arrange for higher $z$ in our model in a number of ways. Watts and Strogatz\cite{WS98} proposed adding bonds to next-nearest or further neighbors on the underlying one-dimensional lattice up to some fixed range which we will call $k$\cite{note2}. In our variation on the model we can also start with such a lattice and then add shortcuts to it. The mean number of shortcuts is then $\phi kL$ and the average coordination number is $z=2k(1+\phi)$. Fig.~\ref{onedim}(b) shows a realization of this model for $k=3$. \begin{figure} \begin{center} \psfig{figure=twodim.ps,width=\columnwidth} \end{center} \caption{(a) An example of a $k=1$ small-world graph with an underlying lattice of dimension $d=2$. (b)~The pattern of bonds around a vertex on the $d=2$ lattice for $k=3$.} \label{twodim} \end{figure} Another way of increasing the coordination number, suggested first by Watts\cite{WattsThesis,Watts99}, is to use an underlying lattice for the model with dimension greater than one. In this paper we will consider networks based on square and (hyper)cubic lattices in $d$ dimensions. We take a lattice of linear dimension $L$, with $L^d$ vertices, nearest-neighbor bonds and periodic boundary conditions, and add shortcuts between randomly chosen pairs of vertices. Such a graph has $\phi dL^d$ shortcuts and an average coordination number $z=2d(1+\phi)$. An example is shown in Fig.~\ref{twodim}(a) for $d=2$. We can also add bonds between next-nearest or further neighbors to such a lattice. The most straightforward generalization of the one-dimensional case is to add bonds along the principal axes of the lattice up to some fixed range $k$, as shown in Fig.~\ref{twodim}(b) for $k=3$. Graphs of this type have $\phi kdL^d$ shortcuts on average and a mean coordination number of $z=2kd(1+\phi)$. Our main interest in this paper is with the properties of the small-world model for small values of the shortcut probability $\phi$. Watts and Strogatz\cite{WS98} found that the model displays many of the characteristics of true random graphs even for $\phi\ll1$, and it seems to be in this regime that the model's properties are most like those of real-world social networks. \section{Length-scales in small-world graphs} \label{length} A fundamental observable property of interest on small-world lattices is the shortest path between two vertices---the number of degrees of separation---measured as the number of bonds traversed to get from one vertex to another, averaged over all pairs of vertices and over all realizations of the randomness in the model. We denote this quantity $\ell$. On ordinary regular lattices $\ell$ scales linearly with the lattice size $L$. On the underlying lattices used in the models described here for instance, it is equal to $\frac14 dL/k$. On true random graphs, in which the probability of connection between any two vertices is the same, $\ell$ is proportional to $\log N/\log z$, where $N$ is the number of vertices on the graph\cite{Bollobas85}. The small-world model interpolates between these extremes, showing linear scaling $\ell\sim L$ for small $\phi$, or on systems small enough that there are very few shortcuts, and logarithmic scaling $\ell\sim\log N = d\log L$ when $\phi$ or $L$ is large enough. In this section and the following one we study the nature of the cross-over between these two regimes, which we refer to as ``large-world'' and ``small-world'' regimes respectively. For simplicity we will work mostly with the case $k=1$, although we will quote results for $k>1$ where they are of interest. When $k=1$ the small-world model has only one independent parameter---the probability $\phi$---and hence can have only one non-trivial length-scale other than the lattice constant of the underlying lattice. This length-scale, which we will denote $\xi$, can be defined in a number of different ways, all definitions being necessarily proportional to one another. One simple way is to define $\xi$ to be the typical distance between the ends of shortcuts on the lattice. In a one-dimensional system with $k=1$, for example, there are on average $\phi L$ shortcuts and therefore $2\phi L$ ends of shortcuts. Since the lattice has $L$ vertices, the average distance between ends of shortcuts is $L/(2\phi L) = 1/(2\phi)$. In fact, it is more convenient for our purposes to define $\xi$ without the factor of $2$ in the denominator, so that $\xi=1/\phi$, or for general $d$ \begin{equation} \xi = {1\over(\phi d)^{1/d}}. \label{defsxi} \end{equation} For $k>1$ the appropriate generalization is\cite{note3} \begin{equation} \xi = {1\over (\phi kd)^{1/d}}. \label{generalxi} \end{equation} As we see, $\xi$ diverges as $\phi\to0$ according to\cite{note4} \begin{equation} \xi \sim \phi^{-\tau}, \label{defstau} \end{equation} where the exponent $\tau$ is \begin{equation} \tau = {1\over d}. \label{valuetau} \end{equation} A number of authors have previously considered a divergence of the kind described by Eq.~\eref{defstau} with $\xi$ defined not as the typical distance between the ends of shortcuts, but as the system size $L$ at which the cross-over from large- to small-world scaling occurs\cite{BA99,Barrat99,NW99,MMP99}. We will shortly argue that in fact the length-scale $\xi$ defined here is precisely equal to this cross-over length, and hence that these two divergences are the same. The quantity $\xi$ plays a role similar to that of the correlation length in an interacting system in standard statistical physics. Its leaves the system with no length-scale other than the lattice spacing, so that at long distances we expect all spatial distributions to be scale-free. This is precisely the behavior one sees in an interacting system undergoing a continuous phase transition, and it is reasonable to regard the small-world model as having a continuous phase transition at this point. Note that the transition is a one-sided one since $\phi$ is a probability and cannot take values less than zero. In this respect the transition is similar to that seen in the one-dimensional Ising model, or in percolation on a one-dimensional lattice. The exponent $\tau$ plays the part of a critical exponent for the system, similar to the correlation length exponent $\nu$ for a thermal phase transition. De Menezes~{\it{}et~al.}\cite{MMP99} have argued that the length-scale $\xi$ can {\em only\/} be defined in terms of the cross-over point between large- and small-world behavior, that there is no definition of $\xi$ which can be made consistent in the limit of large system size. For this reason they argue that the transition at $\phi=0$ should be regarded as first-order rather than continuous. In fact however, the arguments of de~Menezes~{\it{}et~al.}\ show only that one particular definition of $\xi$ is inconsistent; they show that $\xi$ cannot be consistently defined in terms of the mean vertex--vertex distance between vertices in finite regions of infinite small-world graphs. This does not prove that no definition of $\xi$ is consistent in the $L\to\infty$ limit and, as we have demonstrated here, consistent definitions do exist. Thus it seems appropriate to consider the transition at $\phi=0$ to be a continuous one. Barth\'el\'emy and Amaral\cite{BA99} have conjectured on the basis of numerical simulations that $\tau=\frac23$ for $d=1$. As we have shown here, $\tau$ is in fact equal to $1/d$, and specifically $\tau=1$ in one dimension. We have also demonstrated this result previously using a renormalization group~(RG) argument\cite{NW99}, and it has been confirmed by extensive numerical simulations\cite{Barrat99,NW99,MMP99}. The length-scale $\xi$ governs a number of other properties of small-world graphs. First, as mentioned above, it defines the point at which the average vertex--vertex distance $\ell$ crosses over from linear to logarithmic scaling with system size $L$. This statement is necessarily true, since $\xi$ is the only non-trivial length scale in the model, but we can demonstrate it explicitly by noting that the linear scaling regime is the one in which the average number of shortcuts on the lattice is small compared with unity and the logarithmic regime is the one in which it is large\cite{WattsThesis}. The cross-over occurs in the region where the average number of shortcuts is about one, or in other words when $\phi k dL^d=1$. Rearranging for $L$, the cross-over length is \begin{equation} L = {1\over(\phi k d)^{1/d}} = \xi. \end{equation} The length-scale $\xi$ also governs the average number $V(r)$ of neighbors of a given vertex within a neighborhood of radius $r$. The number of vertices in such a neighborhood increases as $r^d$ for $r\ll\xi$ while for $r\gg\xi$ the graph behaves as a random graph and the size of the neighborhood must increase exponentially with some power of $r/\xi$. To derive the specific functional form of $V(r)$ we consider a small-world graph in the limit of infinite $L$. Let $a(r)$ be the surface area of a ``sphere'' of radius $r$ on the underlying lattice of the model, i.e.,~it is the number of points which are exactly $r$ steps away from any vertex. (For $k=1$, $a(r) = 2^d r^{d-1}/\Gamma(d)$ when $r\gg1$.) The volume within a neighborhood of radius $r$ in an infinite system is the sum of $a(r)$ over $r$, plus a contribution of $V(r-r')$ for every shortcut encountered at a distance $r'$, of which there are on average $2\xi^{-d} a(r')$. Thus $V(r)$ is in general the solution of the equation \begin{equation} V(r) = \sum_{r'=0}^r a(r') [1+2\xi^{-d} V(r-r')]. \end{equation} In one dimension with $k=1$, for example, $a(r)=2$ for all $r$ and, approximating the sum with an integral and then differentiating with respect to $r$, we get \begin{equation} {{\rm d} V\over{\rm d} r} = 2[1 + 2V(r)/\xi], \end{equation} which has the solution \begin{equation} V(r) = \mbox{$\frac12$}\xi({\rm e}^{4r/\xi} - 1). \label{volume} \end{equation} Note that for $r\ll\xi$ this scales as $r$, independent of $\xi$, and for $r\gg\xi$ it grows exponentially, as expected. Eq.~\eref{volume} also implies that the surface area of a sphere of radius $r$ on the graph, which is the derivative of $V(r)$, should be \begin{equation} A(r) = 2{\rm e}^{4r/\xi}. \label{area} \end{equation} These results are easily checked numerically and give us a simple independent measurement of $\xi$ which we can use to confirm our earlier arguments. In Fig.~\ref{converge} we show curves of $A(r)$ from computer simulations of systems with $\phi=0.01$ for values of $L$ equal to powers of two from $128$ up to $131\,072$ (solid lines). The dotted line is Eq.~\eref{area} with $\xi$ taken from Eq.~\eref{defsxi}. The convergence of the simulation results to the predicted exponential form as the system size grows confirms our contention that $\xi$ is well-defined in the limit of large $L$. Fig.~\ref{radii} shows $A(r)$ for $L=100\,000$ for various values of $\phi$. Eq.~\eref{area} implies that the slope of the lines in the limit of small $r$ is $4/\xi$. In the inset we show the values of $\xi$ extracted from fits to the slope as a function of $\phi$ on logarithmic scales, and a straight-line fit to these points gives us an estimate of $\tau=0.99\pm0.01$ for the exponent governing the transition at $\phi=0$ (Eq.~\eref{defstau}). This is in good agreement with our theoretical prediction that $\tau=1$. \begin{figure} \begin{center} \psfig{figure=converge.ps,width=\columnwidth} \end{center} \caption{The mean surface area $A(r)$ of a neighborhood of radius $r$ on a $d=1$ small-world graph with $\phi=0.01$ for $L=128\ldots131\,072$ (solid lines). The measurements are averaged over $1000$ realizations of the system each. The dotted line is the theoretical result for $L=\infty$, Eq.~\eref{area}.} \label{converge} \end{figure} \begin{figure} \begin{center} \psfig{figure=radii.ps,width=\columnwidth} \end{center} \caption{The mean surface area $A(r)$ of a neighborhood of radius $r$ on a $d=1$ small-world graph with $L=100\,000$ for $\phi=10^{-4}\ldots10^{-2}$. The measurements are averaged over $1000$ realizations of the system each. Inset: the value of $\xi$ extracted from the curves in the main figure, as a function of $\phi$. The gradient of the line gives the value of the exponent $\tau$, which is found by a least squares fit (the dotted line) to be $0.99\pm0.01$.} \label{radii} \end{figure} \section{Scaling in small-world graphs} \label{scaling} Given the existence of the single non-trivial length-scale $\xi$ for the small-world model, we can also say how the mean vertex--vertex distance $\ell$ should scale with system size and other parameters near the phase transition. In this regime the dimensionless quantity $\ell/L$ can be a function only of the dimensionless quantity $L/\xi$, since no other dimensionless combinations of variables exist. Thus we can write \begin{equation} \ell = L f(L/\xi), \end{equation} where $f(x)$ is an unknown but universal scaling function. A scaling form similar to this was suggested previously by Barth\'el\'emy and Amaral\cite{BA99} on empirical grounds. Substituting from Eq.~\eref{defsxi}, we then get for the $k=1$ case \begin{equation} \ell = L f(\phi^{1/d} L). \end{equation} (We have absorbed a factor of $d^{1/d}$ into the definition of $f(x)$ here to make it consistent with the definition we used in Ref.~\onlinecite{NW99}.) The usefulness of this equation derives from the fact that the function $f(x)$ contains no dependence on $\phi$ or $L$ other than the explicit dependence introduced through its argument. Its functional form can however change with dimension $d$ and indeed it does. In order to obey the known asymptotic forms of $\ell$ for large and small systems, the scaling function $f(x)$ must satisfy \begin{equation} f(x) \sim {\log x\over x}\qquad\mbox{as $x\to\infty$}, \label{upperlim} \end{equation} and \begin{equation} f(x) \to \mbox{$\frac14$} d\qquad\mbox{as $x\to0$}. \label{lowerlim} \end{equation} When $k>1$, $\ell$ tends to $\frac14 dL/k$ for small values of $L$ and $\xi$ is given by Eq.~\eref{generalxi}, so the appropriate generalization of the scaling form is \begin{equation} \ell = {L\over k} f\bigl((\phi k)^{1/d} L\bigr), \label{scalinglaw} \end{equation} with $f(x)$ taking the same limiting forms~\eref{upperlim} and~\eref{lowerlim}. Previously we derived this scaling form in a more rigorous way using an RG argument\cite{NW99}. We can again test these results numerically by measuring $\ell$ on small-world graphs for various values of $\phi$, $k$ and $L$. Eq.~\eref{scalinglaw} implies that if we plot the results on a graph of $\ell k/L$ against $(\phi k)^{1/d} L$, they should collapse onto a single curve for any given dimension~$d$. In Fig.~\ref{collapse} we have done this for systems based on underlying lattices with $d=1$ for a range of values of $\phi$ and $L$, for $k=1$ and~$5$. As the figure shows, the collapse is excellent. In the inset we show results for $d=2$ with $k=1$, which also collapse nicely onto a single curve. The lower limits of the scaling functions in each case are in good agreement with our theoretical predictions of $\frac14$ for $d=1$ and $\frac12$ for $d=2$. \begin{figure} \begin{center} \psfig{figure=collapse.ps,width=\columnwidth} \end{center} \caption{Data collapse for numerical measurements of the mean vertex--vertex distance on small-world graphs with $d=1$. Circles and squares are results for $k=1$ and $k=5$ respectively for values of $L$ between $128$ and $32\,768$ and values of $\phi$ between $1\times10^{-6}$ and $3\times10^{-2}$. Each point is averaged over 1000 realizations of the randomness. In all cases the errors on the points are smaller than the points themselves. The dashed line is the second-order series approximation with exact coefficients given in Eq.~\eref{series}, while the dot-dashed line is the fifth-order approximation using numerical results for the last three coefficients. The solid line is the third-order Pad\'e approximant, Eqs.~\eref{pade} and~\eref{padesoln}. Inset: data collapse for two-dimensional systems with $k=1$ for values of $L$ from 64 to $1024$ and $\phi$ from $3\times10^{-6}$ up to $1\times10^{-3}$.} \label{collapse} \end{figure} We are not able to solve exactly for the form of the scaling function $f(x)$, but we can express it as a series expansion in powers of $\phi$ as follows. Since the scaling function is universal and has no implicit dependence on $k$, it is adequate to calculate it for the case $k=1$; its form is the same for all other values of $k$. For $k=1$ the probability of having exactly $m$ shortcuts on the graph is \begin{equation} P_m = \biggl({dL^d\atop m}\biggr) \,\phi^m (1-\phi)^{dL^d-m}. \end{equation} Let $\ell_m$ be the mean vertex--vertex distance on a graph with $m$ shortcuts in the limit of large $L$, averaged over all such graphs. Then the mean vertex--vertex distance averaged over all graphs regardless of the number of shortcuts is \begin{equation} \ell = \sum_{m=0}^{dL^d} P_m \ell_m. \label{expansion} \end{equation} Note that in order to calculate $\ell$ up to order $\phi^m$ we only need to know the behavior of the model when it has $m$ or fewer shortcuts. For the $d=1$ case the values of the $\ell_m$ have been calculated up to $m=2$ by Strang and Eriksson\cite{SE99} and are given in Table~\ref{table1}. Substituting these into Eq.~\eref{expansion} and collecting terms in $\phi$, we then find that \begin{equation} {\ell\over L} = \mbox{$\frac14 - \frac{1}{24} \phi L + \frac{11}{1440} \phi^2 L^2 - \frac{11}{1440} \phi^2 L + {\rm O}(\phi^3)$}. \end{equation} The term in $\phi^2 L$ can be dropped when $L$ is large or $\phi$ small, since it is negligible by comparison with at least one of the terms before it. Thus the scaling function is \begin{equation} f(x) = \mbox{$\frac14 - \frac{1}{24} x + \frac{11}{1440} x^2 + {\rm O}(x^3)$}. \label{series} \end{equation} This form is shown as the dotted line in Fig.~\ref{collapse} and agrees well with the numerical calculations for small values of the scaling variable $x$, but deviates badly for large values. \begin{table} \begin{tabular}{cccc} & $m$ & $\ell_m/L$ & \\ \tableline & 0 & $1/4$ & \\ & 1 & $5/24$ & \\ & 2 & $131/720$ & \\ & 3 & $0.1549\pm0.0003$ & \\ & 4 & $0.1365\pm0.0003$ & \\ & 5 & $0.1232\pm0.0003$ & \\ \end{tabular} \null\vspace{3mm} \caption{Average vertex--vertex distances per vertex $\ell_m/L$ on $d=1$ small-world graphs with exactly $m$ shortcuts and $k=1$. Values up to $m=2$ are the exact results of Strang and Eriksson\protect\cite{SE99}. Values for $m=3\ldots5$ are our numerical results.} \label{table1} \end{table} Calculating the exact values of the quantities $\ell_m$ for higher orders is an arduous task and probably does not justify the effort involved. However, we have calculated the values of the $\ell_m$ numerically up to $m=5$ by evaluating the average vertex--vertex distance $\ell$ on graphs which are constrained to have exactly 3, 4 or 5 shortcuts. Performing a Taylor expansion of $\ell/L$ about $L=\infty$, we get \begin{equation} {\ell\over L} = {\ell_m\over L} \Bigl[1 + {c\over L} + {\rm O}\bigl(L^{-2}\bigr)\Bigr], \end{equation} where $c$ is a constant. Thus we can estimate $\ell_m/L$ from the vertical-axis intercept of a plot of $\ell/L$ against $L^{-1}$ for large $L$. The results are shown in Table~\ref{table1}. Calculating higher orders still would be straightforward. Using these values we have evaluated the scaling function $f(x)$ up to fifth order in $x$; the result is shown as the dot--dashed line in Fig.~\ref{collapse}. As we can see the range over which it matches the numerical results is greater than before, but not by much, indicating that the series expansion converges only slowly as extra terms are added. It appears therefore that series expansion would be a poor way of calculating $f(x)$ over the entire range of interest. A much better result can be obtained by using our series expansion coefficients to define a Pad\'e approximant to $f(x)$\cite{GG74,note6}. Since we know that $f(x)$ tends to a constant $f(0)=\frac14 d$ for small $x$ and falls off approximately as $1/x$ for large $x$, the appropriate Pad\'e approximants to use are odd-order approximants where the approximant of order $2n+1$ ($n$ integer) has the form \begin{equation} f(x) = f(0) {A_n(x)\over B_{n+1}(x)}, \end{equation} where $A_n(x)$ and $B_n(x)$ are polynomials in $x$ of degree $n$ with constant term equal to~1. For example, to third order we should use the approximant \begin{equation} f(x) = f(0) {1 + a_1 x\over1 + b_1 x + b_2 x^2}. \label{pade} \end{equation} Expanding about $x=0$ this gives \begin{eqnarray} {f(x)\over f(0)} &=& 1 + (a_1 - b_1) x + (b_1^2 - a_1 b_1 - b_2) x^2\nonumber\\ & & + [(a_1 - b_1) (b_1^2 - b_2) + b_1 b_2] x^3 + {\rm O}(x^4).\nonumber\\ \end{eqnarray} Equating coefficients order by order in $x$ and solving for the $a$'s and $b$'s, we find that \begin{eqnarray} a_1 &=& 1.825\pm0.075,\nonumber\\ \label{padesoln} b_1 &=& 1.991\pm0.075,\\ b_2 &=& 0.301\pm0.012.\nonumber \end{eqnarray} Substituting these back into~\eref{pade} and using the known value of $f(0)$ then gives us our approximation to $f(x)$. This approximation is plotted as the solid line in Fig.~\ref{collapse} and, as the figure shows, is an excellent guide to the value of $f(x)$ over a large range of $x$. In theory it should be possible to calculate the fifth-order Pad\'e approximant using the numerical results in Table~\ref{table1}, although we have not done this here. Substituting $f(x)$ back into the scaling form, Eq.~\eref{scalinglaw}, we can also use the Pad\'e approximant to predict the value of the mean vertex--vertex distance for any values of $\phi$, $k$ and $L$ within the scaling regime. We will make use of this result in the next section to calculate the effective dimension of small-world graphs. \section{Effective dimension} \label{seceffdim} The calculation of the volumes and surface areas of neighborhoods of vertices on small-world graphs in Section~\ref{length} leads us naturally to the consideration of the dimension of these systems. On a regular lattice of dimension $D$, the volume $V(r)$ of a neighborhood of radius $r$ increases in proportion to $r^D$, and hence one can calculate $D$ from\cite{note7} \begin{equation} D = {{\rm d}\log V\over{\rm d}\log r} = {r A(r)\over V(r)}, \label{defsd1} \end{equation} where $A(r)$ is the surface area of the neighborhood, as previously. We can use the same expression to calculate the effective dimension of our small-world graphs. Thus in the case of an underlying lattice of dimension $d=1$, the effective dimension of the graph is \begin{equation} D = {4r\over\xi}\,{{\rm e}^{4r/\xi}\over{\rm e}^{4r/\xi}-1}, \label{effdim} \end{equation} where we have made use of Eqs.~\eref{volume} and~\eref{area}. For $r\ll\xi$ this tends to one, as we would expect, and for $r\gg\xi$ it tends to $4r/\xi$, increasing linearly with the radius of the neighborhood. Thus the effective dimension of a small-world graph depends on the length-scale on which we look at it, in a way reminiscent of the behavior of multifractals\cite{Mandelbrot74,HJKPS86}. This result will become important in Section~\ref{percolation} when we consider site percolation on small-world graphs. In Fig.~\ref{dimension} we show the effective dimension of neighborhoods on a large graph measured in numerical simulations (circles), along with the analytic result, Eq.~\eref{effdim} (solid line). As we can see from the figure, the numerical and analytic results are in good agreement for small radii $r$, but the numerical results fall off sharply for larger $r$. The reason for this is that Eq.~\eref{defsd1} breaks down as $V(r)$ approaches the volume of the entire system; $V(r)$ must tend to $L^d$ in this limit and hence the derivative in~\eref{defsd1} tends to zero. The same effect is also seen if one tries to use Eq.~\eref{defsd1} on ordinary regular lattices of finite size. To characterize the dimension of an entire system therefore, we use another measure of $D$ as follows. \begin{figure} \begin{center} \psfig{figure=dimension.ps,width=\columnwidth} \end{center} \caption{Effective dimension $D$ of small-world graphs. The circles are results for $D$ from numerical calculations on an $L=1\,000\,000$ system with $d=1$, $k=1$ and $\phi=10^{-3}$ using Eq.~\eref{defsd1}. The errors on the points are in all cases smaller than the points themselves. The solid line is Eq.~\eref{effdim}. The squares are calculated from Eq.~\eref{otherd} by numerical differentiation of simulation results for the scaling function $f(x)$ of one-dimensional systems. The dotted line is Eq.~\eref{otherd} evaluated using the third-order Pad\'e approximant to the scaling function derived in Section~\ref{scaling}. Inset: effective dimension from Eq.~\eref{otherd} plotted as a function of the scaling variable $x$. The dotted lines represent the asymptotic forms for large and small $x$ discussed in the text.} \label{dimension} \end{figure} On a regular lattice of finite linear size $\ell$, the number of vertices $N$ scales as $\ell^D$ and hence we can calculate the dimension from \begin{equation} D = {{\rm d}\log N\over{\rm d}\log\ell}. \label{defsd2} \end{equation} We can apply the same formula to the calculation of the effective dimension of small-world graphs putting $N=L^d$, although, since we don't have an analytic solution for $\ell$, we cannot derive an analytic solution for $D$ in this case. On the other hand, if we are in the scaling regime described in Section~\ref{scaling}---the regime in which $\xi\gg1$---then Eq.~\eref{scalinglaw} applies, along with the limiting forms, Eqs.~\eref{upperlim} and~\eref{lowerlim}. Substituting into~\eref{defsd2}, this gives us \begin{equation} {1\over D} = {{\rm d}\log\ell\over{\rm d}\log L^d} = {1\over d} \biggl[1+{{\rm d}\log f(x)\over{\rm d}\log x}\biggr], \label{otherd} \end{equation} where $x = (\phi k)^{1/d} L \propto L/\xi$. In other words $D$ is a universal function of the scaling variable~$x$. We know that $f(x)$ tends to a constant for small $x$ (i.e.,~$\xi\gg L$), so that $D=d$ in this limit, as we would expect. For large $x$ (i.e.,~$\xi\ll L$), Eq.~\eref{upperlim} applies. Substituting into~\eref{otherd} this gives us $D = d\log x$. In the inset of Fig.~\ref{dimension} we show $D$ from numerical calculations as a function of $x$ in one-dimensional systems of a variety of sizes, along with the expected asymptotic forms, which it follows reasonably closely. In the main figure we also show this second measure of $D$ (squares with error bars) as a function of the system radius $\ell$ (with which it should scale linearly for large $\ell$, since $\ell\sim\log x$ for large $x$). As the figure shows, the two measures of effective dimension agree reasonably well. The numerical errors on the first measure, Eq.~\eref{defsd1} are much smaller than those on the second, Eq.~\eref{defsd2} (which is quite hard to calculate numerically), but the second measure is clearly preferable as a measure of the dimension of the entire system, since the first fails badly when $r$ approaches $\ell$. We also show the value of our second measure of dimension calculated using the Pad\'e approximant to $f(x)$ derived in Section~\ref{scaling} (dotted line in the main figure). This agrees well with the numerical evaluation for radii up to about $1000$ and has significantly smaller statistical error, but overestimates $D$ somewhat beyond this point because of inaccuracies in the approximation; the Pad\'e approximant scales as $1/x$ for large values of $x$ rather than $\log x/x$, which means that $D$ will scale as $x$ rather than $\log x$ for large $x$. \section{Percolation} \label{percolation} In the previous sections of this paper we have examined statistical properties of small-world graphs such as typical length-scales, vertex--vertex distances, scaling of volumes and areas, and effective dimension of graphs. These are essentially static properties of the networks; to the extent that small-world graphs mimic social networks, these properties tell us about the static structure of those networks. However, social science also deals with dynamic processes going on within social networks, such as the spread of ideas, information, or diseases. This leads us to the consideration of dynamical models defined on small-world graphs. A small amount of research has already been conducted in this area. Watts\cite{WattsThesis,Watts99}, for instance, has considered the properties of a number of simple dynamical systems defined on small-world graphs, such as networks of coupled oscillators and cellular automata. Barrat and Weigt\cite{BW99} have looked at the properties of the Ising model on small-world graphs and derived a solution for its partition function using the replica trick. Monasson\cite{Monasson99} looked at the spectral properties of the Laplacian operator on small-world graphs, which tells us about the time evolution of a diffusive field on the graph. There is also a moderate body of work in the mathematical and social sciences which, although not directly addressing the small-world model, deals with general issues of information propagation in networks, such as the adoption of innovations\cite{Rogers62,CKM66,Strang91,Valente96}, human epidemiology\cite{SS88,KM88,Logini88}, and the flow of data on the Internet\cite{KW91,HK99}. In this section we discuss the modeling of information or disease propagation specifically on small-world graphs. Suppose for example that the vertices of a small-world graph represent individuals and the bonds between them represent physical contact by which a disease can be spread. The spread of ideas can be similarly modeled; the bonds then represent information connections between individuals which could include letters, telephone calls, or email, as well as physical contacts. The simplest model for the spread of disease is to have the disease spread between neighbors on the graph at a uniform rate, starting from some initial carrier individual. From the results of Section~\ref{scaling} we already know what this will look like. If for example we wish to know how many people in total have contracted a disease, that number is just equal to the number $V(r)$ within some radius $r$ of the initial carrier, where $r$ increases linearly with time. (We assume that no individual can catch the disease twice, which is the case with most common diseases.) Thus, Eq.~\eref{volume} tells us that, for a $d=1$ small-world graph, the number of individuals who have had a particular disease increases exponentially, with a time-constant governed by the typical length-scale $\xi$ of the graph. Since all real-world social networks have a finite number of vertices $N$, this exponential growth is expected to saturate when $V(r)$ reaches $N=L^d$. This is not a particularly startling result; the usual model for the spread of epidemics is the logistic growth model, which shows initial exponential spread followed by saturation. For a disease like influenza, which spreads fast but is self-limiting, the number of people who are ill at any one time should be roughly proportional to the area $A(r)$ of the neighborhood surrounding the initial carrier, with $r$ again increasing linearly in time. This implies that the epidemic should have a single humped form with time, like the curves of $A(r)$ plotted in Fig.~\ref{radii}. Note that the vertical axis in this figure is logarithmic; on linear axes the curves are bell-shaped rather than quadratic. In the context of the spread of information or ideas, similar behavior might be seen in the development of fads. By a fad we mean an idea which is catchy and therefore spreads fast, but which people tire of quickly. Fashions, jokes, toys, or buzzwords might be expected to show popularity profiles over time similar to the curves in Fig.~\ref{radii}. However, for most real diseases (or fads) this is not a very good model of how they spread. For real diseases it is commonly the case that only a certain fraction $p$ of the population is susceptible to the disease. This can be mimicked in our model by placing a two-state variable on each vertex which denotes whether the individual at that vertex is susceptible. The disease then spreads only within the local ``cluster'' of connected susceptible vertices surrounding the initial carrier. One question which we can answer with such a model is how high the density $p$ of susceptible individuals can be before the largest connected cluster covers a significant fraction of the entire network and an epidemic ensues. Mathematically, this is precisely the problem of site percolation on a social network, at least in the case where the susceptible individuals are randomly distributed over the vertices. To the extent that small-world graphs mimic social networks, therefore, it is interesting to look at the percolation problem. The transition corresponds to the point on a regular lattice at which a percolating cluster forms whose size increases with the size $L$ of the lattice for arbitrarily large $L$\cite{SA92}. On random graphs there is a similar transition, marked by the formation of a so-called ``giant component'' of connected vertices\cite{AS92}. On small-world graphs we can calculate approximately the percolation probability $p=p_c$ at which the transition takes place as follows. Consider a $d=1$ small-world graph of the kind pictured in Fig.~\ref{onedim}. For the moment let us ignore the shortcut bonds and consider the percolation properties just of the underlying regular lattice. If we color in a fraction $p$ of the sites on this underlying lattice, the occupied sites will form a number of connected clusters. In order for two adjacent parts of the lattice not to be connected, we must have a series of at least $k$ consecutive unoccupied sites between them. The number $n$ of such series can be calculated as follows. The probability that we have a series of $k$ unoccupied sites starting at a particular site, followed by an occupied one is $p(1-p)^k$. Once we have such a series, the states of the next $k$ sites are fixed and so it is not possible to have another such series for $k$ steps. Thus the number $n$ is given by \begin{equation} n = p(1-p)^k (L-kn). \label{basicn} \end{equation} Rearranging for $n$ we get \begin{equation} n = L\,{p(1-p)^k\over1+kp(1-p)^k}. \label{defsn} \end{equation} For this one-dimensional system, the percolation transition occurs when we have just one break in the chain, i.e.,~when $n=1$. This gives us a $k$th order equation for $p_c$ which is in general not exactly soluble, but we can find its roots numerically if we wish. Now consider what happens when we introduce shortcuts into the graph. The number of breaks $n$, Eq.~\eref{defsn}, is also the number of connected clusters of occupied sites on the underlying lattice. Let us for the moment suppose that the size of each cluster can be approximated by the average cluster size. A number $\phi kL$ of shortcuts are now added to the graph between pairs of vertices chosen uniformly at random. A fraction $p^2$ of these will connect two occupied sites and therefore can connect together two clusters of occupied sites. The problem of when the percolation transition occurs is then precisely that of the formation of a giant component on an ordinary random graph with $n$ vertices. It is known that such a component forms when the mean coordination number of the random graph is one\cite{AS92}, or alternatively, when the number of bonds on the graph is a half the number of vertices. In other words, the transition probability $p_c$ must satisfy \begin{equation} p_c^2\phi kL = \mbox{$\frac12$} L\,{p_c(1-p_c)^k\over1+kp_c(1-p_c)^k}, \end{equation} or \begin{equation} \phi = {(1-p_c)^k\over 2kp_c[1+kp_c(1-p_c)^k]}. \label{solnpc} \end{equation} We have checked this result against numerical calculations. In order to find the value of $p_c$ numerically, we employ a tree-based invasion algorithm similar to the invaded cluster algorithm used to find the percolation point in Ising systems\cite{MCLSC95,BN99}. This algorithm can calculate the entire curve of average cluster size versus $p$ in time which scales as $L\log L$\cite{note8}. We define $p_c$ to be the point at which the average cluster size divided by $L$ rises above a certain threshold. For systems of infinite size the transition is instantaneous and hence the choice of threshold makes no difference to $p_c$, except that $p_c$ can never take a value lower than the threshold itself, since even in a fully connected graph the average cluster size per vertex can be no greater than the fraction $p_c$ of occupied vertices. Thus it makes sense to choose the threshold as low as possible. In real calculations, however, we cannot use an infinitesimal threshold because of finite size effects. For the systems studied here we have found that a threshold of $0.2$ works well. \begin{figure} \begin{center} \psfig{figure=pc.ps,width=\columnwidth} \end{center} \caption{Numerical results for the percolation threshold on $L=10\,000$ small-world graphs with $k=1$ (circles), 2~(squares), and 5~(triangles) as a function of the shortcut density $\phi$. The solid lines are the analytic approximation to the same quantity, Eq.~\eref{solnpc}.} \label{pc} \end{figure} Fig.~\ref{pc} shows the critical probability $p_c$ for systems of size $L=10\,000$ for a range of values of $\phi$ for $k=1$, 2 and 5. The points are the numerical results and the solid lines are Eq.~\eref{solnpc}. As the figure shows the agreement between simulation and theory is good although there are some differences. As $\phi$ approaches one and the value of $p_c$ drops, the two fail to agree because, as mentioned above, $p_c$ cannot take a value lower than the threshold used in its calculation, which was $0.2$ in this case. The results also fail to agree for very low values of $\phi$ where $p_c$ becomes large. This is because Eq.~\eref{defsn} is not a correct expression for the number of clusters on the underlying lattice when $n<1$. This is clear since when there are no breaks in the sequence of connected vertices around the ring it is not also true that there are no connected clusters. In fact there is still one cluster; the equality between number of breaks and number of clusters breaks down at $n=1$. The value of $p$ at which this happens is given by putting $n=1$ in Eq.~\eref{basicn}. Since $p$ is close to one at this point its value is well approximated by \begin{equation} p \simeq 1 - L^{-1/k}, \end{equation} and this is the value at which the curves in Fig.~\ref{pc} should roll off at low $\phi$. For $k=5$ for example, for which the roll-off is most pronounced, this expression gives a value of $p\simeq0.8$, which agrees reasonably well with what we see in the figure. There is also an overall tendency in Fig.~\ref{pc} for our analytic expression to overestimate the value of $p_c$ slightly. This we put down to the approximation we made in the derivation of Eq.~\eref{solnpc} that all clusters of vertices on the underlying lattice can be assumed to have the size of the average cluster. In actual fact, some clusters will be smaller than the average and some larger. Since the shortcuts will connect to clusters with probability proportional to the cluster size, we can expect percolation to set in within the subset of larger-than-average clusters before it would set in if all clusters had the average size. This makes the true value of $p_c$ slightly lower than that given by Eq.~\eref{solnpc}. In general however, the equation gives a good guide to the behavior of the system. We have also examined numerically the behavior of the mean cluster radius $\rho$ for percolation on small-world graphs. The radius of a cluster is defined as the average distance between vertices within the cluster, along the edges of the graph within the cluster. This quantity is small for small values of the percolation probability $p$ and increases with $p$ as the clusters grow larger. When we reach percolation and a giant component forms it reaches a maximum value and then drops as $p$ increases further. The drop happens because the percolating cluster is most filamentary when percolation has only just set in and so paths between vertices are at their longest. With further increases in $p$ the cluster becomes more highly connected and the average shortest path between two vertices decreases. \begin{figure} \begin{center} \psfig{figure=rho.ps,width=\columnwidth} \end{center} \caption{Average cluster radius $\rho$ as a function of the percolation probability $p$ for site percolation on small-world graphs with $k=1$, $\phi=0.1$ and $L$ equal to a power from $512$ up to $16\,384$ (circles, squares, diamonds, upward-pointing triangles, left-pointing triangles and downward-pointing triangles respectively). Each set of points is averaged over 100 realizations of the corresponding graph. Inset: the same data collapsed according to Eq.~\eref{rhoscale} with $\nu=0.59$, $\gamma=1.3$ and $p_c=0.74$.} \label{rho} \end{figure} By analogy with percolation on regular lattices we might expect the average cluster radius for a given value of $\phi$ to satisfy the scaling form\cite{SA92} \begin{equation} \rho = \ell^{\gamma/\nu} \widetilde{\rho}\bigl((p-p_c) \ell^{1/\nu}\bigr), \label{rhoscale} \end{equation} where $\widetilde{\rho}(x)$ is a universal scaling function, $\ell$ is the radius of the entire system and $\gamma$ and $\nu$ are critical exponents. In fact this scaling form is not precisely obeyed by the current system because the exponents $\nu$ and $\gamma$ depend in general on the dimension of the lattice. As we showed in Section~\ref{seceffdim}, the dimension $D$ of a small-world graph depends on the length-scale on which you look at it. Thus the value of $D$ ``felt'' by a cluster of radius $\rho$ will vary with $\rho$, implying that $\nu$ and $\gamma$ will vary both with the percolation probability and with the system size. If we restrict ourselves to a region sufficiently close to the percolation threshold, and to a sufficiently small range of values of $\ell$, then Eq.~\eref{rhoscale} should be approximately correct. \begin{figure} \begin{center} \psfig{figure=nupc.ps,width=\columnwidth} \end{center} \caption{Best fit values of $p_c$ as a function of $1/\nu$. Inset: the values are calculated from the vertical-axis intercept of a plot of the position $p_0$ of the peak of $\rho$ against $\ell^{-1/\nu}$ (see Eq.~\eref{intercept}).} \label{nupc} \end{figure} In Fig.~\ref{rho} we show numerical data for $\rho$ for small-world graphs with $k=1$, $\phi=0.1$ and $L$ equal to a power of two from 512 up to $16\,384$. As we can see, the data show the expected peaked form, with the peak in the region of $p=0.8$, close to the expected position of the percolation transition. In order to perform a scaling collapse of these data we need first to extract a suitable value of $p_c$. We can do this by performing a fit to the positions of the peaks in $\rho$\cite{NB99}. Since the scaling function $\widetilde{\rho}(x)$ is (approximately) universal, the positions of these peaks all occur at the same value of the scaling variable $y = (p-p_c) \ell^{1/\nu}$. Calling this value $y_0$ and the corresponding percolation probability $p_0$, we can rearrange for $p_0$ as a function of $\ell$ to get \begin{equation} p_0 = p_c + y_0 \ell^{-1/\nu}. \label{intercept} \end{equation} Thus if we plot the measured positions $p_0$ as a function of $\ell^{-1/\nu}$, the vertical-axis intercept should give us the corresponding value of $p_c$. We have done this for a single value of $\nu$ in the inset to Fig.~\ref{nupc}, and in the main figure we show the resulting values of $p_c$ as a function of $1/\nu$. If we now perform our scaling collapse, with the restriction that the values of $\nu$ and $p_c$ fall on this line, then the best coincidence of the curves for $\rho$ is obtained when $p_c=0.74$ and $\nu=0.59\pm0.05$---see the inset to Fig.~\ref{rho}. The value of $\gamma$ can be found separately by requiring the heights of the peaks to match up, which gives $\gamma=1.3\pm0.1$. The collapse is noticeably poorer when we include systems of size smaller than $L=512$, and we attribute this not merely to finite size corrections to the scaling form, but also to variation in the values of the exponents $\gamma$ and $\nu$ with the effective dimension of the percolating cluster. The value $p_c=0.74$ is in respectable agreement with the value of $0.82$ from our direct numerical measurements. We note that $\nu$ is expected to tend to $\mbox{$\frac12$}$ in the limit of an infinite-dimensional system. The value $\nu=0.59$ found here therefore confirms our contention that small-world graphs have a high effective dimension even for quite moderate values of $\phi$, and thus are in some sense close to being random graphs. (On a two-dimensional lattice by contrast $\nu=\frac43$.) \section{Conclusions} \label{concs} In this paper we have studied the small-world network model of Watts and Strogatz, which mimics the behavior of networks of social interactions. Small-world graphs consist of a set of vertices joined together in a regular lattice, plus a low density of ``shortcuts'' which link together pairs of vertices chosen at random. We have looked at the scaling properties of small-world graphs and argued that there is only one typical length-scale present other than the fundamental lattice constant, which we denote $\xi$ and which is roughly the typical distance between the ends of shortcuts. We have shown that this length-scale governs the transition of the average vertex--vertex distance on a graph from linear to logarithmic scaling with increasing system size, as well as the rate of growth of the number of vertices in a neighborhood of fixed radius about a given point. We have also shown that the value of $\xi$ diverges on an infinite lattice as the density of shortcuts tends to zero, and therefore that the system possesses a continuous phase transition in this limit. Close to the phase transition, where $\xi$ is large, we have shown that the average vertex--vertex distance on a finite graph obeys a simple scaling form and in any given dimension is a universal function of a single scaling variable which depends on the density of shortcuts, the system size and the average coordination number of the graph. We have calculated the form of the scaling function to fifth order in the shortcut density using a series expansion and to third order using a Pad\'e approximant. We have defined two measures of the effective dimension $D$ of small-world graphs and find that the value of $D$ depends on the scale on which you look at the graph in a manner reminiscent of the behavior of multifractals. Specifically, at length-scales shorter than $\xi$ the dimension of the graph is simply that of the underlying lattice on which it is built, and for length-scales larger than $\xi$ it increases linearly, with a characteristic constant proportional to $\xi$. The value of $D$ increases logarithmically with the number of vertices in the graph. We have checked all of these results by extensive numerical simulation of the model and in all cases we find good agreement between the analytic predictions and the simulation results. In the last part of the paper we have looked at site percolation on small-world graphs as a model of the spread of information or disease in social networks. We have derived an approximate analytic expression for the percolation probability $p_c$ at which a ``giant component'' of connected vertices forms on the graph and shown that this agrees well with numerical simulations. We have also performed extensive numerical measurements of the typical radius of connected clusters on the graph as a function of the percolation probability and shown by performing a scaling collapse that these obey, to a reasonable approximation, the expected scaling form in the vicinity of the percolation transition. The characteristic exponent $\nu$ takes a value close to $\mbox{$\frac12$}$, indicating that, as far as percolation is concerned, the graph's properties are close to those of a random graph. \section*{Acknowledgments} We thank Luis Amaral, Alain Barrat, Marc Barth\'el\'emy, Roman Koteck\'y, Marcio de Menezes, Cris Moore, Cristian Moukarzel, Thadeu Penna, and Steve Strogatz for helpful comments and conversations, and Gilbert Strang and Henrik Eriksson for communicating to us some results from their forthcoming paper. This work was supported in part by the Santa Fe Institute and by funding from the NSF (grant number PHY--9600400), the DOE (grant number DE--FG03--94ER61951), and DARPA (grant number ONR N00014--95--1--0975).
\section{#1} \label{#2}} \newcommand{\Sec}[2] { \section{#1} \label{#2}} \newcommand{\eq}[1] {Eq.~(\ref{#1})} \newcommand{\Th}[1] {Theorem~\ref{#1}} \newcommand{\Lem}[1] {Lemma~\ref{#1}} \newcommand{\Cor}[1] {Cor.~\ref{#1}} \newcommand{\Con}[1] {Con.~\ref{#1}} \newcommand{\Fig}[1] {Fig.~\ref{#1}} \newcommand{\Secref}[1] {\S\ref{#1}} \newcommand{\Tblref}[1] {Table~\ref{#1}} \newcommand{\makev}[2]{\left( \begin{array}{c}{#1}\\{#2}\end{array}\right)} \newcommand{H\'enon }{H\'enon } \newcommand{\rule[-1.2mm]{0mm}{4mm}}{\rule[-1.2mm]{0mm}{4mm}} \newcommand{\per}[1]{(#1)^{\infty}} \newcommand{\homoc}[1]{+^\infty-(#1)-+^\infty} \newcommand{\homo}[1]{(#1)} \newcommand{\Oe} {{\cal O}(\epsilon)} \newcommand{k_{\rm sn}}{k_{\rm sn}} \newcommand{k_{\rm pf}}{k_{\rm pf}} \newcommand{\epsilon_{\rm sn}}{\epsilon_{\rm sn}} \newcommand{\epf}{\epsilon_{\rm pf}} \newcommand{\asn} {\mbox{\rm asn}} \newcommand{\sn} {\mbox{\rm sn}} \newcommand{\sigmaF} {\Sigma_{\cal{F}}} \newcommand{\pf} {\mbox{\rm pf}} \newcommand{\mbox{\rm pd}}{\mbox{\rm pd}} \newcommand{\bif}[2]{#1 \, \{ #2 \} } \newcommand{\ptcbif}[3]{#1 \rightarrow \bif{#2}{#3} } \renewcommand{\qed} {$\Box$ \\} \newcommand{\qedd} {$\Box$ } \newcommand{\Ws} {{W}^{s}} \newcommand{\Wu} {{W}^{u}} \newcommand{\mbox{{\Huge .}}}{\mbox{{\Huge .}}} \newcommand{\hootnote}[1]{\footnote{#1}} \newcommand{\mathop{\rm Tr}}{\mathop{\rm Tr}} \newcommand{\mathop{\rm sign}}{\mathop{\rm sign}} \newcommand{\mathop{\rm type}}{\mathop{\rm type}} \newcommand{\mathop{\rm fix}}{\mathop{\rm fix}} \newcommand{{\rm trans}}{{\rm trans}} \newcommand{\mathbf{G}}{\mathbf{G}} \newcommand{\mathbf{p}}{\mathbf{p}} \newcommand{\bs} {\mathbf{s}} \newcommand{\bt} {\mathbf{t}} \newcommand{\bT} {\mathbf{T}} \newcommand{\bH} {\mathbf{H}} \newcommand{\bz} {\mathbf{z}} \newcommand{\mbox{diag}}{\mbox{diag}} \newcommand{\id} {\mbox{id}} \newcommand{\mR} {\mathbb{R}} \newcommand{\mZ} {\mathbb{Z}} \newcommand{\marginpar{\rule[-2ex]{2mm}{3ex}{\tiny }}}{\marginpar{\rule[-2ex]{2mm}{3ex}{\tiny }}} \newcommand{\hdnote}[1]{\footnote{{\bf Q}: #1}} \newcommand{\balpha} {\beta} \pagestyle{headings} \begin{document} \title{Homoclinic Bifurcations for the H\'enon Map} \author{D. Sterling, H.R. Dullin \& J.D. Meiss \thanks{Useful conversations with R. Easton and B. Peckham are gratefully acknowledged. DS was supported in part by NSF traineeship grant number DMS-9208685, JDM was supported in part by NSF grant number DMS-9623216 and HRD was supported by DFG grant number Du 302}\\ Department of Applied Mathematics\\ University of Colorado\\ Boulder, CO 80309 } \maketitle \begin{abstract} Chaotic dynamics can be effectively studied by continuation from an anti-integrable limit. We use this limit to assign global symbols to orbits and use continuation from the limit to study their bifurcations. We find a bound on the parameter range for which the H\'enon map exhibits a complete binary horseshoe as well as a subshift of finite type. We classify homoclinic bifurcations, and study those for the area preserving case in detail. Simple forcing relations between homoclinic orbits are established. We show that a symmetry of the map gives rise to constraints on certain sequences of homoclinic bifurcations. Our numerical studies also identify the bifurcations that bound intervals on which the topological entropy is apparently constant. \\ \textbf{AMS classification scheme numbers}: 58F05, 58F03, 58C15 \end{abstract} \Sec{Introduction}{introsec} The problem of determining the sequence of bifurcations that result in the creation of a Smale horseshoe is an interesting one \cite{Easton86, Grassberger89, Davis91}. In this paper we use a continuation technique based on an ``anti-integrable'' (AI) limit \cite{Aubry91} to study some of these bifurcations from the opposite side, that is, as bifurcations that destroy the horseshoe. As a simple example, we study the family of H\'enon maps \cite{Henon68,Henon76} \begin{equation} \label{henmap} \makev {{x'}}{{y'}} = \makev{ {y-k+x^2} }{ {-bx}} . \end{equation} Apart from the fact that the H\'enon maps are the simplest, non-trivial maps of the plane, they are of more general interest as well, since vector fields in the neighborhood of certain codimension-two homoclinic bifurcations can be reduced to H\'enon-like maps \cite{Naudot96,CHS96}. As we recall in \Secref{sec:ai}, the AI limit for this map is essentially $k \rightarrow \infty$. In order to represent this with finite parameters, we need only rescale the map, letting \[ z = \epsilon x \;, \mbox{ where } \; \epsilon = \frac{1}{\sqrt{k}} \; . \] As was shown by Devaney and Nitecki \cite{Devaney79}, the H\'enon map has a hyperbolic horseshoe when \begin{equation} \label{krange} k > (1+|b|)^{2} \frac{5+2\sqrt5}{4} \; . \end{equation} The H\'enon map has at most $2^{n}$ \cite{Moser60} periodic points of period $n$, and when the map has a hyperbolic horseshoe, all these orbits exist and are easily identified by their symbolic labels. We showed earlier \cite{Sterling98a} that a contraction mapping argument implies there is a one-to-one correspondence between orbits in the AI limit and bounded orbits of the H\'enon map in precisely the range \eq{krange}. In \Secref{sec:subshift} we show that if we consider a particular subset of orbits, this bound can be increased. Moreover, in \Secref{sec:horse}, we present results of numerical investigations for all $b$ that give what we believe are optimal bounds. In general, the existence of an anti-integrable limit leads to a natural symbolic characterization of orbits---for the H\'enon map this is the same as the horseshoe coding. We use this coding, and as discussed in \Secref {sec:numerics}, a predictor-corrector continuation method \cite{Seydel}, to give each orbit a {\it global code}. That is, we label an orbit with the AI code, and use this designation for the family of orbits until it collides with a family with a different code. For the H\'enon map, this gives a map from the bounded orbits of the map to sequences of symbols $\bs \in 2^{\mZ}$ modulo cyclic permutations, providing only that every orbit can be {\em smoothly} connected to the AI limit. This is the working hypothesis for our numerical method, even though we know that it is probably not true in general. It is certainly valid when the hyperbolic horseshoe exists. We give an example of a periodic orbit not smoothly connected to the AI limit in the dissipative case. This illustrates a general anti-monotonicity result \cite{Yorke92}, stating that when $b \not = -1,0,1$ the map generically creates but also destroys orbits when $k$ is increased, so that the topological entropy is not necessarily monotone. We extend this anti-monotonicity result to the area preserving case by a quite different argument concerning the vanishing of twist in the neighborhood of the period tripling bifurcations in a separate paper \cite{DMS99}. Even though the area preserving case exhibits anti-monotonicity, we still conjecture that there are no isolated bubbles in the bifurcation diagram, i.e.\ that every orbit is continuously connected to the AI limit. Our global code contrasts with other methods for constructing symbolic dynamics for maps, which rely on some attempt to obtain a generating partition \cite{Cvitanovic88, Grassberger89, Hao91, Hansen92, Christiansen95b}. These methods rely on somewhat ad hoc techniques for constructing the partition, especially when there exist elliptic orbits. Our method gives a natural partition that is smoothly connected to the horseshoe, though it does rely on our working hypothesis In our computations of the H\'enon map, we observe that the horseshoe destroying bifurcation appears to be a homoclinic saddle-node bifurcation when the map is orientation preserving, and a heteroclinic saddle-node when it is not. In \Secref{sec:homo} we study in detail the homoclinic orbits of the area preserving H\'enon map, and show how the AI code directly gives other properties of the orbits, such as their ``type,'' ``transition time,'' and ``Poincar\'e signature.'' For an area preserving map, the destruction of a horseshoe by a homoclinic bifurcation gives rise generically to elliptic orbits. Specifically, if $f$ is a $ C^{1}$ area preserving diffeomorphism with a homoclinic tangency at $x$ then for any neighborhood $U$ of $x$, there is an area preserving diffeomorphism $C^{1}$-close to $f$ that has an elliptic periodic point in $U$ \cite{New77}. Much more is known about the behavior of area-contracting maps near a homoclinic tangency. Gavrilov and Silnikov proved that if a $C^{3}$ map has a quadratic homoclinic tangency at a parameter $k^{*}$ then there exists a sequence of parameter values $k_{n} \rightarrow k^{*}$ such that at $k_{n}$ there is a saddle-node bifurcation creating orbits of period $n$ \cite{Gav72,Gav73}; because one of the created orbits is a sink, this is called a {\it cascade of sinks}. Robinson extended these results to the real analytic case where the intersection is created degenerately \cite{Rob83}. In our computations we will find a similar cascade of saddle-node bifurcations for the area preserving H\'enon map---this gives a sequence of elliptic orbits that limit on the homoclinic bifurcation. Thus the destruction of the horseshoe is associated with the creation of the first stable ``island.'' The ordering on the invariant manifolds poses severe restrictions on the possible bifurcations. In \Secref{sec:hobs} we use it to prove which homoclinic bifurcation of the hyperbolic fixed point is the first one. We observe that the forcing relations between homoclinic orbits up to type 6 is essentially like the unimodal ordering of one dimensional maps. Generically a homoclinic bifurcation corresponds to a quadratic tangency of the stable and unstable manifolds---a ``homoclinic saddle-node bifurcation.'' There are two more generic bifurcations in maps with a symmetry: a homoclinic pitchfork when the manifolds exhibit a cubic tangency, and a simultaneous pair of asymmetric saddle node bifurcations. In \Secref{sec:pitchfork} we show that a symmetric homoclinic bifurcation forbids certain other bifurcations to occur after it, leading to a natural mechanism to create homoclinic pitchfork bifurcations or asymmetric saddle node pairs. We observe all three of these bifurcations for the area preserving H\'enon map, which has a time-reversal symmetry. Davis, MacKay, and Sannami \cite{Davis91} conjectured that there are intervals of $k$ below the horseshoe for which the H\'enon map is a hyperbolic Markov shift. They also identified the Markov partitions for these cases. Their conjecture was based on computing all the periodic orbits up to a period $20$ using the technique of Biham and Wenzel \cite{Biham89, Biham90}. In \Secref{sec:entropy}, we confirm their computations with our continuation technique and extend them to period $24$---an order of magnitude more orbits. Moreover, we identify the bifurcations responsible for the creation and destruction of these apparently hyperbolic intervals; as befits with the theme of this paper, they are homoclinic bifurcations. \Sec{Anti-Integrable Limit}{sec:ai} Dynamics in discrete time can be represented by a relation, $F(x,x') = 0$ where $x$ and $x'$ are points in some manifold. Normally, we can explicitly solve for $x' = f(x)$, giving a map on the manifold, with orbits defined by sequences $x_{t}=f(x_{t-1})$. Suppose, however, that $F$ depends upon a parameter $\epsilon$, in such a way that this is not always possible; for example, $F(x,x') = \epsilon G(x,x') + H(x)$. In this case the implicit equation $F=0$ can no longer be solved for $x'$ when $\epsilon = 0$; instead ``orbits'' correspond to arbitrary sequences of points, $x_{t}$ that are zeros of $H$---the dynamics are not deterministic. In this case we say that $\epsilon=0$ corresponds to an {\it anti-integrable (AI)} limit of the map $f$ \cite{Aubry95}. If the derivative of $H$ is nonsingular, then a straightforward implicit function argument can be used to show that the AI orbits can be continued for $\epsilon \neq 0$ to orbits of the map $f$ \cite{Aubry91, MacKay92}. An AI limit with this property is called {\it nondegenerate}. For example, consider the H\'enon map \eq{henmap}. Denoting points on an orbit by a sequence $x_{t}$, $t \in \mZ$, we can rewrite \eq{henmap} as a second order difference equation \[ x_{t+1} + bx_{t-1} + k -x_t^2 = 0 \;. \] Introducing the scaled coordinate $z = \epsilon x$ and choosing $\epsilon = k^{-1/2}$ gives an implicit map in the variable $z$ with parameter $\epsilon$ \begin{equation} \label{impmap} \epsilon(z_{t+1} + bz_{t-1}) + 1 -z_t^2 = 0 \;. \end{equation} With this choice of $\epsilon$, we can study only the range $ 0 < k < \infty$; however, one could redefine $\epsilon$ to shift this range.\footnote{ For example, choosing $\hat\epsilon = (k+\delta)^{-1/2}$, maps positive values of $\hat\epsilon$ to the range $-\delta < k < \infty$. Our numerical routines typically use $\delta =1$ so that we can cover the entire parameter range where there are bounded orbits. In the text we always use $\delta = 0$. } A period $n$ orbit of the H\'enon map is given by a sequence $z_0,z_1,\ldots,z_{n-1}$ that satisfies \eq{impmap}, together with the condition that $z_{t+n}=z_{t}$. The corresponding family of periodic orbits is denoted by $\bz(\epsilon)$. At the AI limit, the map \eq{impmap} reduces to \[ z_t^2 = 1 \;. \] Thus ``orbits'' in this limit are arbitrary sequences of $\pm1$, which we abbreviate with $+$ and $-$. Let $\Sigma$ denote the space of such sequences \begin{equation} \Sigma \equiv \{-,+\}^{\mZ} = \{ \bs: s_{t}\in\{-,+\} \,, t \in \mZ \} \,. \end{equation} For ease of notation we denote the corresponding sequence of $\{+1,-1\} \in \mR^\mZ$ by the same symbol $\bs$. Hence every sequence $\bs \in \Sigma$ is an orbit $\bs \in \mR^\mZ$ at the anti-integrable limit, and each of these can be continued to an orbit of the H\'enon map for small enough $\epsilon$ \cite{Aubry90,MacKay92}. Previously we gave an explicit upper bound on $\epsilon$ for the existence of orbits for every symbol sequence \cite{Sterling98a}: \begin{teo} \label{thrm:exist} For every symbol sequence $\bs \in \Sigma$ there exists a unique orbit $\bz(\epsilon)$ of the H\'enon map \eq{impmap}, such that $\bz(0) = \bs$ providing \begin{equation}\label{epsmax} |\epsilon|(1+|b|) < 2\sqrt{1-2/\sqrt{5}} \approx 0.649839 \,. \end{equation} \end{teo} The basic idea of the proof of this theorem is as follows \cite{Sterling98a}. Let $B_{M}$ be the closed ball of radius $M$ around the point $\bs \in \Sigma$, \begin{equation}\label{BMdef} B_{M}(\bs) = \{{\bz} : ||\bz - \bs||_{\infty} \le M \} \;. \end{equation} For each symbol sequence $\bs \in \Sigma$ and small enough $M$, define a map $\bT:B_M \rightarrow B_M$ by \begin{equation}\label{Tdef} T_i(\bz) \equiv s_i\sqrt{1 + \epsilon(z_{i+1} + bz_{i-1})} \;, \end{equation} then the corresponding H\'enon map orbit $\bz(\epsilon)$ is a fixed point of $\bT$. The conclusion of \Th{thrm:exist} follows from finding the maximum value of $\epsilon$ for which there is an $n$ such that $\bT^n$ is a contraction mapping (i.e., $\bT^n:B_M \rightarrow B_M \quad \mbox{and} \quad ||D\bT^n|| < 1$). The fact that $\bT$ is a contraction implies that there are no bifurcations in the range \eq{epsmax}. This is the statement of \begin{cor} There are no bifurcations in the H\'enon map when $\epsilon$ and $b$ are in the range given in \Th{thrm:exist}. \end{cor} {\bf Proof:} Denote the system of equations (\ref{impmap}) by $\bH(\bz,\epsilon) = 0$. This infinite continuation problem has a unique solution $z(\epsilon)$ if the inverse of $D_{z} \bH$ is bounded. The $t$-th component of $\bT$ is related to the $t$-th component of $\bH$ by \[ H_t = T_t^2 - z_t^2. \] Differentiating this at the fixed point that exists according to \Th{thrm:exist} gives \[ D_{z} \bH = \mbox{diag}(2z_t) ( D\bT - \id ). \] The operator $D_{z} \bH$ has bounded inverse because $z_t$ is bounded away from zero by \eq{BMdef} and the inverse of $D\bT-\id$ exists because $||D\bT|| < 1$. \qed This result can be extended to imply that the invariant set is uniformly hyperbolic for the case $b=1$ when the operator $D_{z}\bH$ is symmetric\cite{ABM92}. In \Secref{sec:horse} we will use numerical continuation to estimate the parameters at which the first bifurcation occurs, giving an improvement in this bound, albeit a numerical one. It is interesting that Devaney and Nitecki \cite{Devaney79} obtained precisely the same bound, \eq{epsmax}, for the parameter domain in which the non-wandering set of the H\'enon map is a hyperbolic horseshoe. Nevertheless the AI continuation argument has has two advantages over the geometrical arguments of Devaney and Nitecki. First, it easily generalizes to higher dimensions allowing one to compute parameter bounds for the existence of horseshoes in higher dimensional maps \cite{Sterling98b}. Second, it allows us to easily bound the parameter range for which certain subsets of orbits exist (i.e. the parameter range where the map is conjugate to a subshift of finite type). We present such a bound for a subshift of finite type in \Secref{sec:subshift}. \Sec{Numerical Method}{sec:numerics} In this section we formulate the problem of following H\'enon map orbits away from the anti-integrable limit as a classical continuation problem \cite{Keller}. A period $n$ orbit family of the H\'enon map with coordinates given by $\bz(\epsilon)$ is a zero of the function $\mathbf{G} : \mR^n\times \mR \rightarrow \mR^n$ whose $t^{\rm th}$ component is given by the left hand side of \eq{impmap}. The zeros of $G$ are generically smooth curves in $\mR^n\times \mR$ defined by the continuation problem \[ \label{contsystem} \left\{\begin{array}{lcl} \mathbf{G}(\bz,\epsilon) & = & {\bf 0} \\ \bz(0) & = & \bs \end{array} \right. \;. \] Practically the continuation is always started at the AI limit. The curve might either extend to some maximal $\epsilon_{\rm max}$ and return the the limit or just continue indefinitely. Since for the H\'enon map there are no orbits for $k<-1$ we expect that all the curves do return to the AI limit. This is a standard continuation problem, which we solve using a predictor-corrector method with a linear tangent predictor and a Newton's method corrector. For numerical linear algebra we use the Meschach library \cite{MESCHACH}. The algorithm incorporates an adaptive step size control with bisection backtracking if the corrector fails to converge. The algorithm terminates when a user-specified value of $\epsilon$ is reached or, when the tangent direction is not uniquely defined. The process of continuing a sequence of orbits is trivially parallelizable since the operations performed on each orbit are completely independent of each other. We use a ``divide and conquer'' strategy to spread the total computational effort across several different machines running simultaneously. This is especially advantageous when the number of orbits continued reaches into the millions. Continuation methods are based on the assumption that the orbits of interest are actually connected to the limit at which the continuation starts. Since the H\'enon map does not have an integrable limit, the natural starting point would be to continue all the orbits that bifurcate off the fixed points. But this would only yield a small fraction of all orbits: many of them are born in saddle-node bifurcations that are not connected to any other orbit. Therefore the AI limit is a much better limit from which to continue. However, the same general restriction applies, i.e.~only orbits that are connected to the limit (or their parents, grandparents etc.) can be found. We formulate this central hypothesis as the ``no-bubble-conjecture'': \begin{con} [No Bubbles] For the area preserving H\'enon map every orbit is (at least) continuously connected to the anti-integrable limit. \end{con} We could only hope for continuous connection in $\mR^n\times \mR$ because the branches corresponding to the children in a bifurcation are in general not smoothly connected to the parent. Our numerics currently does not perform any branch-switching from parents to children. Therefore, in practice, we are actually using the working hypothesis: ``Every orbit of the H\'enon map can be {\em smoothly} connected to the anti-integrable limit.'' Unfortunately our working hypothesis is not true in general. For $b \not = -1,0,1$ it has been shown \cite{Yorke92} that periodic orbits are both created and destroyed when the map parameter $k$ is increased (the authors of \cite{Yorke92} call this ``antimonotonicity''). Moreover, in \cite{DMS99} we show that this conclusion holds for the cases $b=\pm 1$ as well. Consequently the topological entropy is not necessarily a monotone increasing function, as it is for the logistic map, $b=0$. In the following we will elucidate the relation between antimonotonicity, our working hypothesis and the no-bubble conjecture. Even though we have antimonotonicity whenever $b\not = 0$, this does not readily imply that our working hypothesis is false. In particular the smallest period orbit that is antimonotonic when $b=1$ is of period $10$, and it is still smoothly connected with the AI limit \cite{DMS99}. However, orbits that bifurcate from this one may violate the working hypothesis.\footnote{ This orbit is created in a $3/10$ rotational bifurcation of the elliptic fixed point, and it initially moves towards smaller $k$ values. While it is traveling in the ``wrong'' direction, the elliptic $3/10$ orbit has a winding number that does not exceed $1/5000$; therefore orbits that bifurcate from it are of period $50000$ or higher. We suspect that these would be orbits that are not smoothly connected to the AI limit, and therefore violate our working hypothesis. Since they bifurcate off the $3/10$ orbit which in turn is smoothly connected, they are, however, at least continuously connected to the AI limit, so that the no-bubble conjecture could still be true. } The worst possible case from the point of view of continuation is an orbit that neither smoothly nor continuously connects to the limit, i.e. an ``isolated bubble.'' Note that in the area preserving case, one orbit of this type implies an infinite number of them because it must be born in a saddle node bifurcation and the stable orbit of the created pair generically passes through an infinite number of rational winding numbers. In order to constitute a violation of our conjecture, none of these orbits would be allowed to reach the AI limit; otherwise the original orbit would be continuously connected. In the dissipative case, the lowest period example we were able to find of a periodic orbit that is not smoothly connected occurs for $b=-0.46$, where at $k\approx 1.0346$ the period $8$ orbit with sequence $\per{-^5+-+}$ has a period doubling bifurcation that creates a period 16 orbit that is not smoothly connected to the AI limit. Since this is the start of a period doubling sequence resulting in a what appears to be a strange attractor at a slightly smaller value of $k$, we expect there are many saddle-node bifurcations in this region that create orbits as $k$ decreases that are (presumably) not connected to the AI limit. This is reason to believe that the ``no bubbles'' conjecture is false when $|b| < 1$. In any case, continuation from the anti-integrable limit has the advantage that at the beginning point all orbits exist, and they all continue nondegenerately. \Sec{Symbolic Dynamics}{sec:symb} In this section we introduce some notation for symbol sequences and bifurcations. For simplicity we concentrate mostly on the area preserving case, $b = 1$, though many results apply generally. Orbits in the anti-integrable limit are bi-infinite sequences $\bs \in \Sigma$. When it is needed, we will indicate the current time along an orbit using a ``$\mbox{{\Huge .}}$'', so that $\bs = \ldots s_{-2}s_{-1} \mbox{{\Huge .}} s_{0}s_{1}s_{2}\ldots$. The dynamics on $\bs \in \Sigma$ are given by the shift map, $\sigma: \Sigma \rightarrow \Sigma$ defined as \[ \sigma (\ldots s_{-1} \mbox{{\Huge .}} s_{0}s_{1}s_{2}\ldots) = \ldots s_{-1}s_{0} \mbox{{\Huge .}} s_{1}s_{2}\ldots \] An orbit of the symbolic dynamics is periodic if the sequence $\bs$ is periodic. We will denote an orbit of least period $n$ by the string of $n$ symbols and a superscript $\infty$ to represent repetition: \[ \per{s_{0}s_{1}\ldots s_{n-1}} = \ldots s_{n-2} s_{n-1} \mbox{{\Huge .}} s_{0} s_{1} \ldots s_{n-1}s_0\ldots \] Of course any cyclic permutation of a periodic orbit gives another point on the same orbit. Trivially, the map $\sigma$ has two fixed points, $\per{+}$ and $\per{-}$, and these correspond to the two fixed points of the H\'enon map. These are born in a saddle-node bifurcation at $k = -(1+b)^{2}/4$, which we denote by \[ \bif{\sn}{ \per{+} ,\per{-}} \;. \] We denote bifurcations with the general template \[ \ptcbif{parent}{type}{children} \;, \] where $parent$ refers to the orbit that is undergoing the bifurcation, if any, and $type$ is one of $\sn$, $\pf$, $\mbox{\rm pd}$, or $m/n$, corresponding to a saddle-node, pitchfork, period doubling, or rotational bifurcation, respectively. The set of orbits created in the bifurcation is listed as the $children$. When the stability of these differ, we adopt the convention that the unstable child is listed first, and the stable one second. When $b=1$ the fixed points of the H\'enon map are located at \[ z_{\pm} = \pm \sqrt{1 +\epsilon^{2}}+\epsilon = \epsilon \left( 1 \pm \sqrt{1+k} \right) \; . \] The stability of a period $n$ orbit of an area preserving map $f$ is conveniently classified by the ``residue'' defined as \[ R = \frac14 \left(2 - \mathop{\rm Tr}(Df^{n}) \right) \;, \] so that an orbit is hyperbolic if it has negative residue, elliptic when $0<R<1$ and is reflection hyperbolic when $R > 1$ \cite{Meiss92}. The residues of the fixed points are \begin{equation} \label{resfixed} R_{\pm} = \mp \frac{1}{2\epsilon} \sqrt{1+\epsilon^{2}} = \mp \frac12 \sqrt{1+k} \;, \end{equation} so the sign of the symbol is opposite to the sign of the residue of the fixed point. Thus the orbit $\per{+}$ is always hyperbolic, while the orbit $\per{-}$ is reflection hyperbolic for small $\epsilon$, or large $k$, but becomes elliptic at $\epsilon = 1/\sqrt3$, or $k=3$. \begin{table}[tb] \centering \begin{tabular}{c|c|cc|c} Parent & Type & Child & Child & $k$-Value \\ \hline & $\sn$ & $\per{-}$ & $\per{+}$ & $-1$ \\ \hline $\per{-}$ & $\mbox{\rm pd}$ & & $\per{+-}$ & $3$ \\ \hline & $\sn$ & $\per{-+-}$ & $\per{+-+}$ & $ 1$ \\ $\per{-}$ & $1/3$ & $\per{-+-}$ & & $ \frac54$ \\ \hline $\per{-}$ & $1/4$ & $\per{+--+}$ & $\per{+-++}$ & $ 0$ \\ $\per{+-}$& $\mbox{\rm pd}$& & $\per{-+--}$ & $ 4$ \\ \hline $\per{-}$ & $1/5$ & $\per{-+++-}$ & $\per{++-++}$ & $\frac{7-5\sqrt5}{8}$ \\ $\per{-}$ & $2/5$ & $\per{--+--}$ & $\per{-+-+-}$ & $\frac{7+5\sqrt5}{8}$ \\ & $\sn$ & $\per{+-+-+}$ & $\per{+---+}$ & $5.5517014^\dagger$ \\ \hline $\per{-}$ & $1/6$ & $\per{-+^{4}-}$ & $\per{++-+^{3}}$ & $-\frac34$ \\ $\per{+-+}$& $\mbox{\rm pd}$& & $\per{+-^{4}+}$ & $\frac54$ \\ $\per{+-^4+}$ & $\pf$ & $\per{++-+--}$ & $\per{--+-++}$ & $3$ \\ & $\sn$ & $\per{--+-+-}$ & $\per{--+-^3}$ & $3.7016569^\ddagger$ \\ $\per{+-}$ & $1/3$ & $\per{--+-+-}$ & & $\frac{15}{4}$ \\ & $\sn$ & $\per{+-+^{3}-}$ & $\per{--+^{3}-}$ & $5.6793695^\ddagger$ \\ \hline \multicolumn{5}{l}{$^\dagger \, 16 k^5-108 k^4+105 k^3+27 k^2-97k-47$}\\ \multicolumn{5}{l}{$^\ddagger \, 16 k^6-136 k^5+213 k^4+220 k^3+126 k^2+108 k+81$}\\ \end{tabular} \caption{Periodic orbits of the H\'enon map up to period 6 and their bifurcations when $b=1$. In the ``type'' column, ``$\sn$'' indicates a saddle-node bifurcation, ``$\pf$'' a pitchfork bifurcation, and ``$\mbox{\rm pd}$'' a period doubling bifurcation. A rotational bifurcation is denoted by $m/n$, referring to the winding number of the parent at the bifurcation. For $1/3$ the child is not created in the bifurcation, it exists before and after the bifurcation. If there are two children, the one listed in the first column has negative residue just after birth (except for the $\pf$ case). The real roots of the polynomials in the last rows give exact bifurcation values for the three approximations shown.} \label{tbl:orbits} \end{table} For $b=1$ the sequence $\per{+-}$ corresponds to the period two orbit \[ \per{+-} \,:\, (z_{0},z_{1}) = (\sqrt{1-3\epsilon^{2}}-\epsilon ,-\sqrt{1-3\epsilon^{2}}-\epsilon) \;. \] This orbit exists only for $\epsilon < 1/\sqrt3$, and is created by a period doubling of the elliptic fixed point (when $R_{-}=1$). We denote this bifurcation by \[ \ptcbif{\per{-}}{\mbox{\rm pd}}{ \per{+-} } \;. \] Similarly there are two period three orbits, \begin{eqnarray*} \per{--+} &:& (z_0, z_1, z_2) = (-\sqrt{1-\epsilon^{2}}, -\sqrt{1-\epsilon^{2}},\sqrt{1-\epsilon^{2}}-\epsilon) \\ \per{-++} &:& (z_0, z_1, z_2) = (-\sqrt{1-\epsilon^{2}}-\epsilon, \sqrt{1-\epsilon^{2}}, \sqrt{1-\epsilon^{2}}) \;. \end{eqnarray*} These are created in a saddle-node bifurcation at $k=\epsilon=1$; \[ \bif{\sn}{ \per{--+} ,\per{-++} } \;. \] We list the low period orbits and their bifurcation values in \Tblref{tbl:orbits}. Another class of bifurcations shown in the table are rotational bifurcations. A rotational bifurcation occurs when the winding number of an elliptic orbit becomes $\omega = m/n$; we denote such bifurcations by the winding number of the parent orbit. For example the birth of orbits with winding number $1/n$ at the fixed point $\per{-}$ is denoted \begin{equation} \label{rotorbits} \ptcbif{\per{-}}{1/n}{\per{--+^{n-2}},\per{-++^{n-2}}} \;. \end{equation} This particular rotational bifurcation occurs when the multipliers of the fixed point are $e^{i2\pi\omega}$ or using \eq{resfixed}, when $k$ is given by \begin{equation} \label{Romega} k_{\omega}= \cos({2\pi \omega})(\cos({2\pi \omega})-2) \;. \end{equation} We have empirically identified the symbol sequences for rotational bifurcations, and will present the general symbolic formula for these and for rotational ``island around island'' orbits in \cite{Sterling98b}. The residue of any periodic orbit of a Lagrangian system is easily computed from the matrix $M$ formed from the second variation of the action \cite{MacKay83}. For a period $n$ orbit of the H\'enon map this formula gives: \[ R(z(\epsilon)) = -\frac14\frac{\det(M)}{\epsilon^{n}} \;, \] where $M$ is the periodic tridiagonal matrix with elements \[ M_{t,t-1}= -b\epsilon \,, \quad M_{t,t} = 2z_t(\epsilon) \,, \quad M_{t,t+1} = -\epsilon \;. \] As we approach the anti-integrable limit, $\bz(\epsilon) \rightarrow \bs$ as $\epsilon \rightarrow 0$ and $M$ approaches the diagonal matrix $\mbox{Diag}(\-2 s_{i})$. Thus we see that the residue becomes infinite at the anti-integrable limit and its sign is given by $- \prod_{t=0}^{n-1}s_{t}$. Hence, \begin{equation} \label{residue} \mathop{\rm sign}(R(\bs)) = -(-1)^{j} \;, \end{equation} where $j$ is the number of minus signs in the symbol sequence $\bs$. \Sec{A Subshift of Finite Type}{sec:subshift} In this section, we extend \Th{thrm:exist} to the case of a subshift of finite type. In particular, the biggest restriction in the proof of the theorem arises from the fact that the lower bound on the operator $\bT$ given in \eq{Tdef} is weakest when the signs $s_{i+1}=s_{i-1}=-1$ for positive $b$. We can improve the bound by restricting the set of admissible symbol sequences to forbid this particular case. The shift map restricted to this subspace is a subshift of finite type with the forbidden set $\cal{F}=\{-+-,-\,-\,-\}$; that is, we define the shift space \[ \sigmaF = \Sigma \setminus \{ \bs : \exists \; t \in \mZ \mbox{ such that } s_{t-1}=s_{t+1}=-\} \; . \] This subshift can be easily described as a subshift on $2$-blocks represented by $\{--,-+,+-,++\}^{\mZ}$. The subshift on the two-block space is represented by a vertex graph with the state transition matrix \cite{LM95} \[ S = \left(\begin{array}{cccc} 0 & 0 & 1 & 0 \\ 1 & 0 & 1 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 1 & 0 & 1 \end{array} \right) \;, \] which indicates by $S_{ij} = 1$ an allowed transitions from state $j$ to state $i$. The two zeros $S_{11}$ and $S_{32}$ come from the forbidden sequence in $\sigmaF$, the remaining are obtained because successive two-blocks overlap in one symbol, i.e.,\ $(s_{t-1}s_{t})$ has to be followed by $(s_{t}s_{t+1})$. The number of fixed points of period $n$ for the subshift is given by \[ \mathop{\rm Tr}(S^{n}) = \gamma^{n}+(1-\gamma)^{n} +2 (-1)^{n/2}(n-1 \mod 2) \;, \] where $\gamma = (1+\sqrt{5})/2$ is the golden mean. Thus the topological entropy for $\sigmaF$ is $ \ln \gamma$. The number of distinct periodic orbits can be obtained from the trace formula by subtracting the number of periodic orbits for all factors of $n$ and then dividing by the number of cyclic permutations, $n$. For comparison with the full shift and with the numerical results below, we give a list of these in \Tblref{tbl:periods}. For example there are a total of $1,465,020$ periodic points of the full shift with period $n \le 24$, while there are only $12,216$ in the subshift $\sigmaF$. {\small \begin{table}[tbp] \centering \begin{tabular}{r|r|r} Period & \multicolumn{1}{c}{$\Sigma$}&\multicolumn{1}{c}{$\sigmaF$} \\ \hline 1 & 2 & 1 \\ 2 & 1 & 0 \\ 3 & 2 & 1 \\ 4 & 3 & 2 \\ 5 & 6 & 2 \\ 6 & 9 & 2 \\ 7 & 18 & 4 \\ 8 & 30 & 5 \\ 9 & 56 & 8 \\ 10 & 99 & 11\\ 11 & 186 & 18\\ 12 & 335 & 25\\ \end{tabular} \hspace*{5pt} \begin{tabular}{r|r|r} Period & \multicolumn{1}{c}{$\Sigma$}&\multicolumn{1}{c}{$\sigmaF$} \\ \hline 13 & 630 & 40\\ 14 & 1161 & 58\\ 15 & 2182 & 90\\ 16 & 4080 & 135 \\ 17 & 7710 & 210 \\ 18 & 14532 & 316 \\ 19 & 27594 & 492 \\ 20& 52377 & 750 \\ 21& 99858 & 1164 \\ 22& 190557 & 1791 \\ 23& 364722 & 2786 \\ 24& 698870 & 4305 \\ \end{tabular} \caption{Number of orbits with minimal period $n$ of the 2-shift and the subshift $\sigmaF$.} \label{tbl:periods} \end{table} } When $b$ is non-negative, orbits with symbol sequences in the subspace $\sigmaF$ can be shown to persist longer than a general orbit: \begin{teo}[Existence and Uniqueness of $\sigmaF$ orbits] \label{thrm:seq} Suppose $0 \le b \le 1$. For every symbol sequence $\bs \in \sigmaF$ there exists a unique orbit $\bz(\epsilon)$ of the H\'enon map \eq{henmap} such that $\bz(0) \equiv \bs$ providing $0 \le \epsilon < \epsilon_{\rm max}$, where \begin{equation} \epsilon_{\rm max} \equiv \frac{2}{1+b}\sqrt{\frac{-b^2+2b+5 - 2\sqrt{5+4b}}{(1-b)(5-b)}} \;. \end{equation} \end{teo} \noindent This theorem follows from the same argument that gave \Th{thrm:exist} with only minor modifications. We summarize the changes in the argument in the following discussion. {\bf Proof:} When $0 \le b \le 1$, $\bs \in \sigmaF$, and $\bz \in B_{M}$, we can bound the norm of iterates of $\bT$ in \eq{Tdef} using the inequalities \[ \alpha_{k} \le ||\bT^{k}(z) ||_{\infty} \le \beta_{k} \;, \] where the coefficients $\alpha_{k}$ and $\beta_{k}$ are determined by the recursions \begin{eqnarray*} \beta_{k} &=& \sqrt{1 +\epsilon(1+b)\beta_{k-1}} \;, \\ \alpha_{k}&=& \sqrt{1 +\epsilon (b\alpha_{n-1} - \beta_{n-1})} \;, \end{eqnarray*} with $\beta_{0}= 1+M$ and $\alpha_{0}= 1-M$. The sequence $\beta_k$ is the same as that in \cite{Sterling98a}; it has the unique attracting fixed point \[ \beta_\infty =\frac12 \left\{\epsilon(1+b) + \sqrt{\epsilon^2(1+b)^2 + 4}\right\} \;. \] Since the recursion for $\{\alpha_n\}$ depends on $\beta$, but not vice versa, the coupled system also has a unique attracting fixed point, which is given by $(\alpha_{\infty},\beta_{\infty})$ with \[ \alpha_\infty = \frac12 \left\{\epsilon b + \sqrt{\epsilon^2b^2 +4(1-\epsilon\beta_\infty)}\right\} \;. \] This implies that for large enough $n$, $\bT^{n}$ maps the ball $B_{1-\alpha_{\infty}}$ into itself. The map $\bT$ is a contraction map on this ball providing $||D\bT^k||_{\infty} < 1 $. This leads to the same bound as that in \cite{Sterling98a}, namely: \[ \epsilon_{\rm max}(1+b) < 2\alpha_\infty\;. \] After some simplification, this inequality yields the formula for $\epsilon_{\rm max}$. The condition that the operator $\bT$ be real, $\epsilon < 1/(\beta_\infty - b \alpha_\infty)$, is satisfied whenever the map is a contraction. \qed Similar arguments for negative $b$ lead to bounds for a subshift with the forbidden subsequence $+\ast -$ (where $\ast$ is any symbol). However, this subshift is not of much interest, since there are only three periodic orbits in it. \Sec{Horseshoe Boundary}{sec:horse} Theorem \ref{thrm:exist} provides an analytical bound on the parameter range for which the H\'enon map has a hyperbolic horseshoe. This bound corresponds to the dark grey region in \Fig{fig:horsebound}. According to \Th{thrm:seq} the subshift $\sigmaF$ exists, in addition, in the lighter shaded area in the figure. This bound is valid only for $b \ge 0$, and meets the former at $b=0$. Here, we use our continuation method to estimate the boundary of existence of the horseshoe by following all orbits up to period 24 from the anti-integrable limit. In order for the numerical boundary to be valid we only need to assume that we can extrapolate from 24 to $\infty$. Since there are at most $2^n$ periodic points of period $n$ in the H\'enon map \cite{Moser60}, we know that we are not missing any orbits because we have them all at the AI limit. This is no longer true after the first bifurcation because orbits that have disappeared might reappear for smaller $k$. \Epsfig{Henon_Horse.eps}{tbp} {First bifurcations for the H\'enon Map. The dark shaded region represents \Th{thrm:exist} and the lighter that of \Th{thrm:seq}. The curves represent the numerical results for the first orbits destroyed up to period $24$. Bounds for the subshift $\sigmaF$ are indicated with a triangle symbol.} {fig:horsebound}{3.5in} To construct a numerical approximation for the boundaries, we first generate all symbol sequences for orbits of periods up to $24$. Then, for fixed $b$, we numerically continue each orbit in $\epsilon$ away from the anti-integrable limit and monitor its multipliers to detect bifurcations. For each $b$ we record the smallest value of $\epsilon$ at which a bifurcation occurs. The resulting numerical bounds in \Fig{fig:horsebound} are shown as solid curves. The numerical bound for the full shift is similar in shape to the analytical one, but shifted to the right in $\epsilon$. While the analytical bound is symmetric under $b \rightarrow -b$, the numerical results are not. For example the first bifurcation at $b=1$ occurs for $\epsilon \approx 0.41888$, while at $b=-1$ it occurs for $\epsilon \approx 0.40167$. In the logistic limit ($b=0$), \eq{henmap} reduces to the logistic map, \begin{equation}\label{logistic} z_{t+1}= \frac{1}{\epsilon}(z_{t}^{2}-1) \;, \end{equation} for which the first bifurcation occurs at $\epsilon = 1/\sqrt2$, where the orbit of the critical point becomes bounded. When $b$ is positive, the symbol sequences for the first pair of orbits destroyed up to period $24$ extrapolate to orbits that are homoclinic to the fixed point $\per{+}$; we conjecture that these are the first orbits destroyed as $\epsilon$ increases from $0$: \begin{con} \label{con:firstbif} For positive $b$, the first bifurcation as $\epsilon$ increases from $0$ corresponds to the homoclinic saddle-node bifurcation \begin{equation} \label{firstbif} \bif{\sn}{+^\infty-(+)-+^\infty, +^\infty-(-)-+^\infty} \;. \end{equation} \end{con} The parenthesis in the middle enclose the ``core'' of the homoclinic orbit, see the next section. A theorem of Smillie \cite{Smillie97} implies that the first bifurcation destroying the H\'enon horseshoe must be a quadratic homoclinic tangency for {\it some} orbit. Our observations imply that it is a homoclinic bifurcation of $\per{+}$. When $b<0$, however, the most natural description of the first bifurcation is as a heteroclinic tangency, which leads to \begin{con} For negative $b$, the first bifurcation as $\epsilon$ increases from $0$ corresponds to the heteroclinic saddle-node bifurcation \[ \bif{\sn}{ -^\infty+(-)-+^\infty , -^\infty+(+)-+^\infty } \;. \] \end{con} This does not contradict Smillie's theorem, as there are many homoclinic bifurcations that accumulate on this heteroclinic bifurcation. For example, for each $m$ the orbits $(-^{m}++-+^m)^{\infty}-+(-^{m}++-+^m)^{\infty}$, are homoclinic to the periodic orbit $(-^{m}++-+^m)^\infty$, and the bifurcation points of these homoclinic orbits limit on that of the heteroclinic orbits as $m \rightarrow \infty$. Filtering the symbol sequences to choose only those in the subshift $\sigmaF$, we can use the same numerical data previously described to find the first bifurcation amongst the orbits in $\sigmaF$. This gives the solid curve marked with triangles in \Fig{fig:horsebound}. This curve has a qualitatively different shape than the analytical bound. For reference we indicate in \Fig{fig:horsebound} the point $k=1.4$, and $b=-0.3$, corresponding to the much studied H\'enon attractor. We also sketch the parameter range ($b$ small enough, $1<k<2$) for which the theorem of Benedicks and Carleson \cite{Benedicks91} implies that the H\'enon map has a transitive attractor with positive Lyapunov exponent. Note that the numerical horseshoe boundary does {\em not} depend on the no bubble conjecture. This is so because it is known that there can be at most $2^n$ periodic points of period $n$ \cite{Moser60}. Since we follow all of them up to period 24 there can be no other orbits up to that period. In other words, orbits first have to be destroyed before they can be reborn. \Sec{Homoclinic Orbits}{sec:homo} In this section we use the symbolic dynamics to classify orbits of the H\'enon map that are homoclinic to the hyperbolic fixed point $p = \per{+}$ and study their bifurcations. We begin with some general terminology, referring to the H\'enon map as an example. Let $f$ be an orientation preserving map\footnote { The orientation reversing case could be included by considering $f^{2}$, since its manifolds have the same geometry as those of $f$. } of the plane with hyperbolic fixed point $p$. The stable and unstable manifolds of $p$ are denoted by $\Wu$ and $\Ws$, and a closed segment of such a manifold between two points $\alpha$ and $\beta$ by $\Wu[\alpha,\beta]$. We use a parenthesis to denote an open endpoint of a segment. A segment that extends to the fixed point, e.g. $\Wu(p,\alpha]$, is called an {\it initial segment} of the manifold. The set of homoclinic orbits is the set of intersections $\Ws \cap \Wu$. A point $\alpha$ is on a {\it primary} (or principal) homoclinic orbit if the two initial segments to $\alpha$ touch only at $\alpha$, i.e., \[ \Wu(p,\alpha] \cap \Ws(p,\alpha] = \{\alpha\} \;. \] Thus the initial segments to a primary homoclinic orbit define a Jordan curve; we call the interior of this curve a resonance zone. More generally, a {\it resonance zone} is a region bounded by alternating initial segments of stable and unstable manifolds \cite{MacKay87,Easton98}. For example, in \Fig{fig:tan} we sketch the left-going branches of the manifolds from $p = \per{+}$ for the area preserving H\'enon map. There are precisely two primary homoclinic orbits; in the figure, we label points on these orbits with $\alpha$ and $\zeta$. We choose to use $\alpha$ to construct the resonance zone. \Epsfig{tangency.eps}{tbp} {Stable and unstable manifolds for the H\'enon map at $k=5$ and $b=1$ shown in $(z,z')$ coordinates.} {fig:tan}{3in} The stable manifold is divided into two invariant branches by the fixed point. An ordering is defined on each branch of $\Ws$, so that $\beta <_{s} \gamma$ for two points on a branch of $\Ws$ if $\beta$ is nearer to $p$ along $\Ws$ than $\gamma$, i.e., $\beta \in \Ws(p,\gamma)$. We similarly define an ordering $<_{u}$ on each branch of $\Wu$. A segment of a manifold from a point to its iterate, $\Ws(\beta,f(\beta)]$, is called a {\it fundamental segment} \cite{Easton98}. The union of the iterates of a fundamental segment is the entire branch of the manifold that contains $\beta$. Moreover, since the iterates are disjoint, every homoclinic orbit on this branch must have precisely one point on each fundamental segment. For the H\'enon map, we focus on the left-going branches of $\Ws$ and $\Wu$ and the fundamental segments between $f^{-1}(\zeta)$ and $\zeta$. These also form the boundaries of the incoming and exit sets for the resonance zone defined by $\alpha$ \cite{Meiss97}. The exterior halves of these segments, $\Ws(\alpha,\zeta)$ and $\Wu(f^{-1}(\zeta),\alpha)$, contain no homoclinic points since orbits on these segments are unbounded, so it is sufficient to look for homoclinic points on the interior halves, \begin{eqnarray} U &\equiv& \Wu[\alpha,\zeta] \;, \nonumber \\ S &\equiv& \Ws[f^{-1}(\zeta),\alpha] \;.\label{SandU} \end{eqnarray} Every homoclinic orbit must have exactly one point on both $S$ and $U$. Homoclinic orbits can be classified in a number of ways. The {\it type} \cite{Easton98}, of a homoclinic point $\beta$ is\footnote { Our definition of the type differs from Easton's slightly, to comply with his definition that type 1 is equivalent to the horseshoe. Rom-Kedar \cite{Rom92} uses the term {\it Birkhoff signature} instead of type. } \[ \mathop{\rm type}(\beta) = \sup \{j \ge 0 : \Ws(p,f^{j}(\beta)] \cap \Wu(p,\beta] \ne \emptyset \} \;; \] i.e., the number of iterates for which the stable initial segment to $f^j(\beta)$ intersects with the unstable initial segment to $\beta$. The type of a homoclinic point is invariant along its orbit. Primary homoclinic points have type 0. Homoclinic orbits on particular branches of $\Ws$ and $\Wu$ can also be classified by their {\it transition time}. In general this is defined relative to a choice of a primary homoclinic point, $\zeta$ and the fundamental segments $\Wu(f^{-1}(\zeta),\zeta]$ and $\Ws(f^{-1}(\zeta),\zeta]$. Any homoclinic orbit on these branches has exactly one point, $\beta$, on the unstable segment. The transition time is the number of iterates required for $\beta \in \Wu(f^{-1}(\zeta),\zeta]$ to reach the stable segment: \[ t_{\rm trans}(\beta) = k \mbox{ if } f^{k}(\beta) \in \Ws(f^{-1}(\zeta),\zeta] \] The value of the transition time depends upon the choice of fundamental segments, so it is not as basic a property as the type. In the simplest case, the transition time is easily related to the type of the orbit \cite{Easton86}: \begin{lem} \label{typelem} Assume there are exactly two primary homoclinic orbits, $\zeta$ and $\alpha$, and the segments $S$ and $U$ defined in \eq{SandU} contain all of the homoclinic orbits. Then for each homoclinic point in $\beta \in U$, $t_{\rm trans}(\beta) = \mathop{\rm type}(\beta)$. \end{lem} {\bf Proof:} If $\beta \in U$ is of type $t$, then by definition $\Ws(p,f^{t}(\beta)] \cap \Wu(p,\beta] \ne \emptyset$. Now since $\alpha <_{u} \beta <_{u} \zeta$ and $W^{s}(p,\zeta) \cap W^{u}(p,\zeta) = \emptyset$, this implies that $\alpha <_{s} f^{t}(\beta)$. However, $\Ws(p,f^{t+1}(\beta)] \cap \Wu(p,\beta] = \emptyset$, which means that $f^{t+1}(\beta) <_{s} \alpha$, but there are no homoclinic points on $\Ws(\zeta,\alpha)$, so actually $f^{t+1}(\beta) <_{s} \zeta$. Now $S$ contains every homoclinic point that reaches $W^{s}(p,\zeta)$ in one iteration, so $f^{t}(\beta) \in S$. \qed Each homoclinic orbit has a Poincar\'e signature that determines the direction of crossing of $\Wu$ and $\Ws$ at points on the orbit. We define the signature to be $+1$ if, looking along the unstable manifold in the direction of motion, the stable manifold crosses the unstable from the left to the right side. Crossings in the opposite direction have signature $-1$. If the manifolds do not cross but only touch (a topologically even intersection), the signature is defined to be zero. Since the map is orientation preserving, the signature is invariant along an orbit. Thus in \Fig{fig:tan} $\alpha$ and $\zeta$ have signatures $-1$ and $+1$, respectively. The signature of a particular homoclinic orbit is typically not preserved in a bifurcation, but the total signature of the bifurcating orbits must be the same on each side of the bifurcation value. For example a saddle-node bifurcation creates a zero signature orbit that splits into one positive and one negative signature orbit. For the H\'enon map, the AI symbol sequence can be used for the classification of homoclinic orbits. It is easy to construct homoclinic and heteroclinic orbits using the symbolic dynamics: an orbit heteroclinic from a periodic orbit $\per{s}$ to a periodic orbit $\per{s'}$ has a symbol sequence that begins with a head sequence $\per{s}$ and ends with a tail sequence $\per{s'}$ with some arbitrary, finite symbol sequence separating the head and tail. For example, the simplest orbits homoclinic to $p = \per{+}$ are the primary homoclinic orbits: \begin{eqnarray} \zeta &=& +^{\infty} - \mbox{{\Huge .}} +^{\infty} \;, \nonumber \\ \alpha &=& +^{\infty} - \mbox{{\Huge .}} - +^{\infty} \;, \end{eqnarray} corresponding to those we labeled in \Fig{fig:tan}. These symbol sequences arise because as $\epsilon \rightarrow 0$ the point $\alpha$ moves to the point $ -\mbox{{\Huge .}}-$, while $\zeta$ moves to $-\mbox{{\Huge .}}+$ and $f^{-1}(\zeta)$ to $+\mbox{{\Huge .}}-$. All other orbits homoclinic to $p$ can be written in the form $+^{\infty}-(s) - +^{\infty}$, where $s$, the {\it core}, is any finite sequence---thus there is a one-to-one correspondence between finite symbol sequences and potential homoclinic orbits (all of which exist in the AI limit). This implies, for example, that near the anti-integrable limit there are $2^{k}$ homoclinic orbits with core length $k$. We will often denote an orbit homoclinic to $p$ simply by writing the core in parenthesis, $(s)$. Note that a given core $(s)$ is not equivalent to any core with the same $s$ cyclically permuted. The classification of homoclinic orbits by their symbol sequence can be used to compute other invariants. To determine the type of an orbit, we simply note that the AI symbols give exactly the same coding for an orbit as the standard symbolic coding for the horseshoe. This implies that the point $+^{\infty}-\mbox{{\Huge .}} (s)-+^{\infty}$ corresponds to a phase point on $U$, and the point $+^{\infty}-(s)\mbox{{\Huge .}} -+^{\infty}$ is on $S$, thus \begin{lem} \label{transitAIL} The transition time of the homoclinic orbit close to the AI limit is given by the length of the core sequence. \end{lem} \noindent For example, the homoclinic orbit $+^{\infty}-(-\,-+)-+^{\infty}$ has the core sequence $(-\,-+)$, and therefore has transition time $3$. Similarly the signature of a homoclinic orbit in the horseshoe is given by simply counting the number of $-$ signs in the core sequence. \begin{lem} The signature of a homoclinic orbit with core $(s)$ close to the AI limit is given by $-(-1)^{j}$ where $j$ is the number of $-$ signs in $s$. \end{lem} \noindent Thus the orbit $(-\,-+)$ has signature $-1$.\footnote {\eq{residue} implies that the signature is the same as the limiting sign of the residue of periodic orbits that approximate the homoclinic orbit.} We will see that, when $b=1$, some homoclinic orbits undergo pitchfork bifurcations, which change their signature, so this rule is not valid for all parameter values. \Epsfig {ordering.eps}{ptb} {Ordering of homoclinic orbits of type 1, 2 and 3 at $k=6.25$. The enlargement on the left shows the core sequences for the $14$ homoclinic orbits up to type $3$ The type 0 orbits $\alpha$ and $\zeta$ are not listed.} {fig:homorder}{4in} The positions of the homoclinic orbits on $U$ for orbits of type 1, 2 and 3, labeled by their core sequences, are shown in \Fig{fig:homorder}. The order of the homoclinic orbits on the segments $S$ and $U$ close to the AI limit must be the same as the corresponding ordering of homoclinic points in the complete horseshoe by continuity. This ordering is equivalent to that of the logistic map, \eq{logistic}, for the orbits forward asymptotic to the $+$ fixed point. This gives an easy way to compute the ordering, see \Fig{fig:logistic}. In the logistic limit, all of the sequences forward asymptotic to the fixed point are destroyed when the orbit of the critical point becomes bounded at $\epsilon = 1/\sqrt2$. The fixed point $\mbox{{\Huge .}} +^{\infty}$ has a single preimage, which is $\alpha = \mbox{{\Huge .}} -+^{\infty}$. Every other orbit that is forward asymptotic to $\mbox{{\Huge .}} +^{\infty}$ has the form $\mbox{{\Huge .}} (s)-+^{\infty}$. \Epsfig {logistic.eps}{tb} {Ordering of pre-periodic points for the $+$ fixed point of the logistic limit of the H\'enon map. The symbols are determined by the itinerary of the orbit relative to the critical point at $z=0$.} {fig:logistic}{4in} In the area preserving case the ordering of the orbits along $S$ is equivalent to that on $U$ upon time reversal. Thus a type $t$ point $+^{\infty}-(s_{1}s_{2}\ldots s_{t}) \mbox{{\Huge .}} -+^{\infty}$ on $S$ is in the same relative position as the point $+^{\infty}-\mbox{{\Huge .}}( s_{t}s_{t-1}\ldots s_{1})-+^{\infty}$ on $U$. Close to the AI limit we always have this ordering on the manifolds, which is just another way of saying that the map is conjugate to the horseshoe map. So long as there are no homoclinic bifurcations, then the orderings $>_u$ and $>_s$ are just given by the usual unimodal ordering as stated in \begin{lem} \label{orderAIL} The ordering $>_u$ on $U$ and $>_s$ on $S$ close to the AI limit is given by \begin{eqnarray*} +^\infty-\mbox{{\Huge .}} e+\dots & >_u & +^\infty-\mbox{{\Huge .}} e-\dots \\ +^\infty-\mbox{{\Huge .}} o+\dots & <_u & +^\infty-\mbox{{\Huge .}} o-\dots \\ \dots+e \mbox{{\Huge .}}-+^\infty & >_s & \dots-e \mbox{{\Huge .}}-+^\infty \\ \dots+o \mbox{{\Huge .}}-+^\infty & <_s & \dots-o \mbox{{\Huge .}}-+^\infty \; , \end{eqnarray*} where $e$ / $o$ are finite sequences with an even / odd number of minus signs, respectively. \end{lem} The ordering shown in \Fig{fig:homorder} and \Fig{fig:logistic} is exactly this one upon appending the ``homoclinic tail'' $-+^\infty$ to the cores. The maximal orbit on $U$ is $\zeta$, corresponding to the tail $\mbox{{\Huge .}} +^\infty$, the minimal orbit is $\alpha$, corresponding to $\mbox{{\Huge .}} -+^\infty$. \Sec{Homoclinic Bifurcations}{sec:hobs} Homoclinic bifurcations are bifurcations between homoclinic orbits. Compared to ordinary bifurcations of periodic orbits they possess additional structure because the invariant manifolds (with their ordering) must be involved in the bifurcation process. To make this explicit, we say that two homoclinic orbits $\beta$ and $\gamma$ are {\it double neighbors} if the segments $\Wu[\beta,\gamma]$ and $\Ws[\beta,\gamma]$ contain no other homoclinic orbits. Three ordered homoclinic points $\beta <_u \gamma <_u \delta$ are {\it triple neighbors} if both $\beta, \gamma$ and $\gamma, \delta$ are double neighbors. An obvious observation with nevertheless important consequences is the ``double neighbor'' lemma: \begin{lem} \label{lem:doubleN} Two homoclinic orbits $\beta$ and $\gamma$ cannot bifurcate unless they are double neighbors. \end{lem} The converse gives a simple forcing relation: before $\beta$ and $\gamma$ can bifurcate any homoclinic orbit on either segment between them must have disappeared. Another consequence is the transition time lemma: \begin{lem} \label{lem:ttime} If two homoclinic orbits $\beta$ and $\gamma$ bifurcate then they must have the same transition time $t_{\rm trans}$. \end{lem} {\bf Proof:} Let $\beta$ and $\gamma$ be neighbors on $U$. If their transition time is different then they are not neighbors on $S$, so they cannot bifurcate. \qed This allows us to extend Lemma~\ref{transitAIL} away from the AI limit, so that one can take the transition time as an adequate replacement of the period: \begin{cor} The transition time of a homoclinic orbit never changes. \end{cor} {\bf Proof:} Since the transition time is an integer it cannot change under smooth deformations. It could only change at bifurcations, but we have just seen that only orbits with the same transition time bifurcate. \qedd \\ Therefore homoclinic bifurcations only take place between double neighbors with the same transition time, i.e., core length. Close to the AI limit the horseshoe is still complete. In this situation it is possible to find all neighbors: \begin{lem} \label{lem:neighbor} Two homoclinic orbits on $U$ are neighbors in the complete horseshoe if and only if they are of the form $+^\infty-\mbox{{\Huge .}}(s+)-+^\infty$ and $+^\infty-\mbox{{\Huge .}}(s-)-+^\infty$. \end{lem} {\bf Proof:} We have to show that there is no homoclinic orbit with core $\delta$ such that $(o+) <_u (\delta) <_u (o-)$ or $(e-) <_u (\delta) <_u (e+)$, where $e = s$ if $s$ has an even number of minus signs or $o = s$ if this number is odd. If the initial string in $\delta$ differs from $s$ then $\delta$ can not be between the sequences $(s-)$ and $(s+)$, therefore $\delta = s\dots$. It is simple to see that \[ \mbox{{\Huge .}} e+\dots \ge_u \mbox{{\Huge .}} e+-+^\infty, \quad \mbox{{\Huge .}} e-\dots \le_u \mbox{{\Huge .}} e--+^\infty \; , \] and similarly \[ \mbox{{\Huge .}} o+\dots \le_u \mbox{{\Huge .}} o+-+^\infty, \quad \mbox{{\Huge .}} o-\dots \ge_u \mbox{{\Huge .}} o--+^\infty \; . \] Since $\delta = s\dots$ it must be of one of the forms on the left hand sides, but then the inequalities show that it is not between $(s+)$ and $(s-)$ hence they must be neighbors. Conversely, suppose we have two neighboring homoclinic orbits (on $U$) $\mbox{{\Huge .}} a$ and $\mbox{{\Huge .}} b$ with $\mbox{{\Huge .}} a <_u \mbox{{\Huge .}} b$. They must differ in at least one symbol so call the first such difference $x$. Their leading common symbols are denoted by $s$, so that $a=s x\alpha$ and $b=s \bar x\beta$ for some sequences $\alpha$ and $\beta$, where $\bar x$ is the opposite symbol to $x$. Applying the ordering relation to the possible combinations of $s$ and $x$ gives either \[ \mbox{{\Huge .}} e - \alpha <_u \mbox{{\Huge .}} (ey)-+^\infty <_u \mbox{{\Huge .}} e + \beta \quad \mbox{or} \quad \mbox{{\Huge .}} o + \alpha <_u \mbox{{\Huge .}} (oy)-+^\infty <_u \mbox{{\Huge .}} o- \beta \;, \] where the choice of the symbol $y$ depends on whether $s$ is even or odd and whether $\alpha$ and $\beta$ are $-+^\infty$. Specifically, choose $y=+$ if either $s=e$ and $\beta \ne -+^\infty$ or $s=o$ and $\alpha \ne -+^\infty$. Choose $y=-$ if either $s=e$ and $\alpha \ne -+^\infty$ or $s=o$ and $\beta \ne -+^\infty$. If neither $\alpha$ nor $\beta$ are $-+^\infty$ either choice for $y$ works. When either $\alpha$, $\beta$, or both differ from $-+^\infty$ we have constructed an orbit $ \mbox{{\Huge .}} (sy)-+^\infty $ which is between $\mbox{{\Huge .}} a$ and $\mbox{{\Huge .}} b$---hence $\mbox{{\Huge .}} a$ and $\mbox{{\Huge .}} b$ are {\it not} neighbors. But this is a contradiction so $\alpha$, $\beta = -+^\infty$. \qed For a bifurcation to occur it is not enough that the orbits be neighbors on $U$, but they must be double neighbors. In the reversible case this almost gives the proof of Conjecture~\ref{con:firstbif}, but here we are working in the smaller class of orbits homoclinic to $p$, the hyperbolic fixed point. So far we did not make use of the reversibility of the map, i.e., the results are valid for all $b$. From this point on we will always only talk about the area preserving case. Note that the ordering is used in a range of parameters before the first bifurcation occurs, so the horseshoe ordering is still valid. \begin{teo} \label{thrm:firstbif} In the area preserving H\'enon map the first homoclinic bifurcation of the invariant manifolds of the fixed point $\per{+}$ is \[ \bif{\sn}{\homoc{+}, \homoc{-}} \;. \] \end{teo} { \bf Proof:} By Lemma~\ref{lem:neighbor} we know that all neighbors on $U$ in the complete horseshoe are of the form $(s\pm)$. For these sequences to be double neighbors they must be neighbors on $S$ as well. By reversibility this is equivalent to the sequence and its reverse being neighbors on $U$. But this only true if $s$ is empty. The only double neighbors in the complete horseshoe are therefore the two orbits $\homoc{+}$ and $\homoc{-}$. Therefore they must bifurcate first. \qed \begin{table}[bt] \centering { \begin{tabular}{c|c|c|c} Orbits & $k_{\rm sn}$ & $k_{\infty}$ & $\delta$ \\ \hline $\per{-*-+^2}$ & 5.5517014388520 & 5.699160106302 & \\ $\per{-*-+^3}$ & 5.6793695105731 & 5.699306445540 & 7.45095 \\ $\per{-*-+^4}$ & 5.6965039879058 & 5.699310669970 & 7.11409 \\ $\per{-*-+^5}$ & 5.6989125149379 & 5.699310783741 & 7.04922 \\ $\per{-*-+^6}$ & 5.6992541878224 & 5.699310786628 & 7.03706 \\ $\per{-*-+^7}$ & 5.6993027411880 & 5.699310786699 & 7.03489 \\ $\per{-*-+^8}$ & 5.6993096429803 & 5.699310786700 & 7.03452 \\ $\per{-*-+^9}$ & 5.6993106241120 & 5.699310786700 & 7.03446 \\ $\per{-*-+^{10}}$ & 5.6993107635871 & & 7.03445 \\ $\per{-*-+^{11}}$ & 5.6993107834145 & & 7.03445 \\ \end{tabular} } \caption{Bifurcations in periodic approximations to the homoclinic type 1 orbit, which is the first orbit destroyed for $b=1$. Here we use a $*$ to denote both $+$ and $-$, giving both orbits involved in the bifurcation. \label{tbl:exit} } \end{table} To approximate a homoclinic orbit, which possesses an infinite number of points in phase space by a periodic orbit with only a finite number of points we require that the Hausdorff distance of these two point sets vanishes as the period approaches infinity. Thus for an orbit homoclinic to $\per{+}$, we study a sequence of approximating periodic orbits with an increasingly long string of $+$ symbols. In particular the rotational orbits given in \eq{rotorbits} converge to $\zeta$ and $\alpha$ in the limit. In \Tblref{tbl:exit} we list the first 11 members of the sequence approximating the transit time $1$ homoclinic orbit, and the corresponding sequence of values, $\epsilon_{\rm sn}$ at which these orbits undergo a saddle-node bifurcation when $b = 1$. These values converge geometrically to the parameter at which the homoclinic orbits bifurcate, and the ratio of successive differences ( a ``Feigenbaum ratio'') is computed in the fourth column of the table. As is known theoretically for $b<1$ \cite{Curry82, Rob83} the convergence rate, $\delta$, approaches $\lambda$, the multiplier of the fixed point $p$. From our data, the convergence rate $\delta$ agrees up to 6 digits with the multiplier \[ \lambda \approx 7.0344478 \] of the fixed point $p$ when $k \approx 5.699310786700$. Thus, our observations indicate that the convergence rate is given by the multiplier in the area preserving case as well, where to our knowledge no proof exists. The third column in the table is the extrapolation for the converged $k$ value, given by Aitken's $\Delta^{2}$ method \[ k_{\infty} = k_n - \frac{\Delta(k_{n})^2}{\Delta^2(k_{n})} \;, \] where $\Delta$ is the forward discrete difference operator. Thus we see that there is a saddle-node bifurcation of the type $1$ homoclinic orbits, \[ \bif{\sn} { +^\infty-(+)-+^\infty, +^\infty-(-)-+^\infty } \;, \] at \[ \epsilon_{\rm sn}(1) \approx 0.418879233367 \quad \mbox{ or }\quad k_{\rm sn}(1) \approx 5.699310786700 \;. \] This also corresponds to the parameter value at which the topological horseshoe for the H\'enon map is destroyed, and is the value in \Fig{fig:horsebound} at $b=1$. Since there is a sequence of saddle-node bifurcations that limit on the homoclinic bifurcation, there are elliptic islands arbitrarily close to the destruction of the horseshoe. This corresponds to an area preserving version of the results of Gavrilov and Silnikov \cite{Gav72,Gav73} and Newhouse \cite{New74, New77}. Since we can in principle follow every finite orbit from the anti-integrable limit we can begin to study the sequence of bifurcations that occur after the horseshoe is destroyed, see \Tblref{tbl:hobis}. For example, the bifurcation diagram for all of the homoclinic orbits of type three or less is sketched in \Fig{fig:bifdiag}. The vertical ordering in this sketch is the same as that on the segment $U$ with $\alpha$ and $\zeta$ shown. The bifurcation diagram is highly influenced by the time-reversal symmetry of the area preserving H\'enon map---we will discuss this symmetry in the next section. As expected from the general theory \cite{Rimmer78}, we observe three kinds of bifurcations: \begin{description} \item[Symmetric saddle-node] bifurcations resulting in the creation of a pair of type $t$ homoclinic orbits with opposite signatures. For example, in \Fig{fig:bifdiag}, the type $3$ orbits with cores $(+++)$ and $(-+-)$ are born in such a saddle-node at $k \approx 0.386$. \item[Pitchfork] bifurcations of type $t$ symmetric homoclinic orbits, creating a pair of type $t$ asymmetric orbits that are related by time reversal. For example, the $(-+-)$ orbit pitchforks at $k \approx 0.720$ creating the orbits $(-++)$ and $(++-)$. A pitchfork bifurcation requires triple neighbors to occur. The parent orbit of a homoclinic pitchfork bifurcations is always created in a symmetric saddle-node bifurcation. Up to type 11 there are are only 9 symmetric saddle-node bifurcations which do {\em not} undergo a homoclinic pitchfork bifurcation on their way to the AI limit. \item[Asymmetric saddle-node] bifurcations creating two symmetry related pairs of asymmetric orbits. This bifurcation first occurs at type $4$. For example the two pairs $\{(-+--), (-+-+)\}$ and $\{(--+-),(+-+-)\}$ are created at $k \approx 5.18$. Generically, asymmetric saddle-node bifurcations require two pairs of double neighbors to occur because of the symmetry. \end{description} The shaded region in \Fig{fig:bifdiag} represents the range of $k$ for which the area preserving H\'enon map exhibits a horseshoe. Along the left edge we label each orbit with its core symbol sequence. \Epsfig{Type3_bif.eps}{tbp} {Sketch of bifurcations in the homoclinic orbits up to type 3 ($b=1$).} {fig:bifdiag}{4in} \begin{table}[btp] \centering \begin{tabular}{c|c|cc|c} Parent & Type & Child & Child & $k$-Value \\ \hline & $\sn$ & $(-++-)$ & $(++++)$ & -0.133474 \\ $(-++-)$ & $\pf$ & $(-+++)$ & $(+++-)$ & -0.044273 \\ & $\sn$ & $(-+-)$ & $(+++)$ & 0.385556 \\ $(-+-)$ & $\pf$ & $(-++)$ & $(++-)$ & 0.719630 \\ & $\sn$ & $(--)$ & $(++) $ & 1.627779 \\ $(--)$ & $\pf$ & $(-+)$ & $(+-) $ & 3.091505 \\ & $\sn$ & $(+--+)$ & $(----)$ & 3.98213640 \\ $(----)$ & $\pf$ & $(---+)$ & $(+---)$ & 3.98213641 \\ & $\sn$ & $(+-+)$ & $(---)$ & 4.706399 \\ $(+-+)$ & $\pf$ & $(--+)$ & $(+--)$ & 4.816792 \\ &$\asn$ & $(-+-*)$ & $(*-+-)$ & 5.188561 \\ &$\asn$ & $(*-++)$ & $(++-*)$ & 5.619922 \\ & $\sn$ & $(+)$ & $(-)$ & 5.699311 \end{tabular} \caption{Homoclinic bifurcations up to core length 4.} \label{tbl:hobis} \end{table} The first type $t$ homoclinic orbits are created by a saddle-node bifurcation when the segment $f^{-t}(S)$ first intersects $U$. We denote this parameter value by $k_{\rm sn}(t)$. This marks the creation of the subset of the incoming lobe of the turnstile with transition time $t$ \cite{Meiss97}. We observe that when $b=1$, this homoclinic saddle-node bifurcation is \[ \bif{\sn}{ (+^{t}) , (-+^{t-2}-) } \quad \mbox{at\ } k_{\rm sn}(t) \;. \] Following this, the orbit $(-+^{t-2}-)$ undergoes a homoclinic pitchfork bifurcation at $k_{\rm pf}(t)$, creating the pair \[ \ptcbif{ (-+^{t-2}-) }{\pf}{ (+^{t-1}-), (-+^{t-1}) } \quad \mbox{at\ } k_{\rm pf}(t) \;. \] However, when $b \ne 1$, the initial symmetric bifurcation and the following symmetry breaking pitchfork are replaced by a pair of nonsymmetric saddle-node bifurcations. In this case the first type $t$ bifurcation is the homoclinic saddle-node \[ \bif{\sn} { (+^{t}), (+^{t-1}-) } \;. \] According to the double neighbor lemma, certain bifurcations cannot occur prior to other homoclinic bifurcations because the corresponding sequences block other sequences from being neighbors. In order to determine which orbits are neighbors even beyond the first bifurcation we make the assumption that the following symbolic ordering conjecture holds: \begin{con}\label{con:orderpersists} The symbolic horseshoe ordering on the invariant manifolds given in Lemma~\ref{orderAIL} persists. \end{con} The ordering relations give a unique construction of the order of the points on $U$ and $S$, and this implies that a schematic construction of the intersections of $f^{-t}(S)$ with $U$ can be constructed solely from a list of which orbits exist at a given parameter value. Such a schematic manifold plot is shown in \Fig{fig:core5mf}, for all homoclinic orbits that exist at $k=5.53$ up to type $5$. We can also construct a schematic bifurcation diagram for homoclinic orbits, as in \Fig{fig:core5bif}, by drawing a horizontal line from $k = \infty$ to the $k$-value at which a particular homoclinic orbit is destroyed---actually we stop the figure at $k=6$, since there are no bifurcations for larger $k$-values. We order the homoclinic orbits vertically according to their unimodal ordering on $U$ as usual. In this bifurcation diagram the vertical connections indicate which orbits eventually do become neighbors and bifurcate. So as to avoid artificially crossing lines, we connect pairs of asymmetric saddle-nodes by lines at the right edge of the figure to indicate that they must bifurcate at the same $k$-value. We say that a bifurcation {\it straddles} the centerline if the pair of orbits involved are on either side of center of the $U$ ordering, or if one of the two pairs of an asymmetric saddle-node straddles the center line. Through type $6$, each symmetric saddle node is followed by a pitchfork bifurcation, just as we observed in \Fig{fig:bifdiag}, with the exception of the very first bifurcation, $\bif{\sn}{(+),(-)}$, which corresponds to the smallest loop straddling the center in the figure. That this is in fact the smallest loop and therefore the first bifurcation is the content of Theorem~\ref{thrm:firstbif}. Moreover, it is remarkable, but perhaps misleading, that through type $6$ every bifurcation straddles the center. Therefore all homoclinic bifurcations up to type 6 are forced by nesting around the center. In particular this means that their unimodal ordering gives the bifurcation ordering, like in unimodal maps. This simple forcing relation is destroyed with the appearance of a symmetric saddle-node without pitchfork of type $7$ (see \Tblref{tbl:gaps}. Also at type $7$, there is an asymmetric saddle-node quadruple which does not straddle the center. Interestingly enough, this is the same bifurcation that marks the upper $k$ endpoint of one of the gaps that we discuss in \Secref{sec:entropy}. \Epsfig{core5mf.eps}{p} {Schematic drawing of $U$ (dashed line) and $f^{-t}(S)$ (solid line) up to type $5$ for $k=5.53$.} {fig:core5mf}{3in} \Epsfig{core5bif.eps}{p} {Bifurcation diagram of homoclinic orbits up to type $5$ ($b=1$). Types 1,2,3 are shown as dotted lines (recall Fig. \ref{fig:bifdiag}); type 4 is dashed; and type 5 is solid.} {fig:core5bif}{4in} It is difficult to visualize the full homoclinic bifurcation diagram for larger type orbits. To do so, we plot only the horizontal lines, to indicate the range of existence of an orbit; this diagram up to transition time 11 is given in \Fig{fig:core11}. The approximate self-similarity in this picture seems to be related to some of the gaps we discuss in \Secref{sec:entropy}, namely those that are related to symmetric saddle-nodes without accompanying pitchforks of type 7, 9 and 11. \EpsfigR{core11.eps}{bt} {Existence plot of homoclinic orbits up to type 11. For each homoclinic orbit a line is drawn from large $k$ to the parameter value where this orbit is destroyed. The vertical position of each line is its formal position on $U$ according to the unimodal ordering.} {fig:core11}{6in} \clearpage \Sec{Symmetric homoclinic bifurcations}{sec:pitchfork} As we mentioned above, the bifurcation diagram of the area preserving H\'enon map is restricted by the fact that the map has a time-reversal symmetry. Here we briefly recall a few well known facts about reversible maps \cite{Lamb98a}, and apply them to the study of homoclinic bifurcations. A map $f$ has a time-reversal symmetry when it is diffeomorphic to its inverse by: \[ Rf = f^{-1} R \;. \] We call the map $R$ a reversor for $f$. Often, as in our case, the reversor is an involution, $R^{2}=I$. Note that each of the maps $f^{t}R$ is also a reversor, in particular, we call $fR$ the complementary reversor to $R$. The fixed set of a reversor \[ \mathop{\rm fix}(R) = \{ x : Rx = x \} \;, \] is of particular interest. For the case when $R$ is an orientation reversing involution of the plane $\mathop{\rm fix}(R)$ is always a curve that goes through infinity, thus dividing the plane into two pieces \cite{MacKay93}. A reversor maps an orbit $\ldots z_{t-1}, z_{t},z_{t+1}\ldots$ of the map onto another orbit $\ldots Rz_{t+1}, Rz_{t}, Rz_{t-1}\ldots$. A symmetric orbit is defined as one that is mapped onto itself by $R$. It is easy to see that any symmetric orbit must have points on $\mathop{\rm fix}(R) \cup \mathop{\rm fix}(fR)$ and conversely. Moreover, if the orbit is not periodic, it has a unique point on one of these fixed sets, and if it is periodic it has exactly two points on the fixed sets \cite{Lamb98b}. Reversible maps need not be area preserving, though the multipliers of an orbit and its symmetric partner must be reciprocals of one another. Application of this to the fixed points gives that the H\'enon map is reversible only when $b = \pm 1$. For a symmetric orbit reversibility implies that the product of the multipliers must be one. For the case $b=\pm 1$ a reversor is $R(x,y) = (-y,-x)$, and a complementary reversor $fR(x,y) = (-x-k+y^{2}, y)$. The fixed curves are \begin{eqnarray*} \mathop{\rm fix}(R) &=& \{(x,y): x=-y\} \;,\\ \mathop{\rm fix}(fR) &=& \{(x,y): x = \frac12 (y^{2}-k) \} \;. \end{eqnarray*} Suppose that $p$ is a symmetric, hyperbolic fixed point of a reversible map. Then, as pointed out by Devaney \cite{Devaney}, the stable and unstable manifolds of the map are related by $R$: \begin{lem} \label{th:symman} Let $\Wu$ and $\Ws$ be the stable and unstable manifolds of a symmetric fixed point $p$. Then $R\Wu(p,\balpha] = \Ws(p,R\balpha]$. \end{lem} {\bf Proof:} By definition, when $\balpha \in \Wu$, then $f^{-t}(\balpha) \rightarrow p$ as $t \rightarrow \infty$. Then $Rf^{-t}(\balpha)= f^{t}(R\balpha) \rightarrow Rp = p$. Thus, $R\balpha \in \Ws$. Since $R$ is a diffeomorphism, $R\Wu(p,\balpha] = \Ws(p,R\balpha]$. \qed \begin{cor} If $\Wu$ intersects the fixed set of a reversor, then the intersection point is homoclinic. \end{cor} Homoclinic orbits of symmetric periodic orbits either come in symmetric pairs or are symmetric, and there must exist symmetric homoclinic orbits: \begin{lem} Let $p$ be a symmetric, hyperbolic fixed point, and $\balpha$ a homoclinic point, and suppose that $R$ is an orientation reversing involution. Then $R\balpha$ is also a homoclinic point. Moreover, there exist symmetric homoclinic points on $\mathop{\rm fix}(R)$ and $\mathop{\rm fix}(fR)$. \end{lem} {\bf Proof:} By Lemma \ref{th:symman}, since $\balpha \in \Ws \cap \Wu$ then $R\balpha \in \Wu \cap \Ws$, so it is homoclinic as well. Since $\mathop{\rm fix}(R)$ divides the plane and $\balpha$ and $R\balpha$ are on opposite sides of this curve, the segment $\Wu[\balpha,R\balpha]$ must cross $\mathop{\rm fix}(R)$, and the crossing point is necessarily homoclinic and symmetric. We can argue similarly for $fR$. \qedd \\ As is well known, pitchfork bifurcations occur with codimension one in maps with a symmetry \cite{Rimmer78}. This occurs for homoclinic bifurcations as well, as was suggested in \cite{Rom95}. We observed such pitchfork bifurcations in \Fig{fig:bifdiag}. A pitchfork typically occurs after a symmetric, type $t > 1$, saddle-node bifurcation creates a ``tip'' of $\Ws$ inside the entry lobe of the turnstile. As this tip grows, one would normally expect it to bend around, as sketched in \Fig{fig:saddle-node}, creating more type $t$ homoclinic points by saddle-node bifurcation. In fact, it is a simple consequence of the linear ordering along $\Wu$ and $\Ws$ combined with reversibility that a single saddle-node bifurcation like that sketched in \Fig{fig:saddle-node} is impossible: \Epsfig{saddle-node.eps}{bt} {Impossibility of the second symmetric bifurcation as described in \Th{pitchfork}. Stable manifolds are shown as solid and unstable manifolds as dashed curves. Reflection of a tangency at $\delta$ gives a tangency at $R\delta$ that is ordered incorrectly.} {fig:saddle-node}{3in} \begin{teo} \label{pitchfork} Suppose that $f$ is an orientation preserving, reversible map, with a symmetric fixed point $p$, and $S$ and $U = RS$ are segments of its stable and unstable manifold bounded by adjacent primary homoclinic orbits. Suppose that a pair of symmetric homoclinic points on $U$, $\beta <_{s} \gamma$ are created in a saddle-node bifurcation. Then it is impossible for there to be a single saddle-node bifurcation as a tangency of $\Ws(\beta,\gamma)$ with either piece of $U \setminus \Wu[\beta,\gamma]$. \end{teo} {\bf Proof:} Since $\beta <_{s} \gamma$, and $R\Ws = \Wu$, we have $R\beta <_{u} R\gamma$. Suppose that $\beta$ and $\gamma$ have transition time $t$. Then $f^{t}(\beta) \in S$, but since $\beta$ is symmetric this point must be the same as $R\beta$. Thus $f^{-t}(R\beta) = \beta$, and similarly for $\gamma$. Since the ordering is preserved by iteration, then $\beta <_{u}\gamma$. Now suppose there is a tangency at a point $\delta = \Ws(\beta,\gamma) \cap (U \setminus \Wu[\beta,\gamma])$, i.e., not between $\beta$ and $\gamma$. We sketch such a configuration in \Fig{fig:saddle-node}. Thus $\beta <_{s} \delta <_{s} \gamma$. By symmetry, $R\beta <_{u} R\delta <_{u} R\gamma$. Since the ordering is preserved by iteration, we have $\beta <_{u} f^{-t}(R\delta) <_{u} \gamma$. Thus $f^{-t}(R\delta)\in \Wu(\beta,\gamma)$ and so this point is not $\delta$ (consequently the orbit of $\delta$ is not symmetric). Since the manifolds are tangent at $\delta$, they are also tangent at $R\delta$ by symmetry. Thus there is a second, simultaneous tangency, on $U$ at $f^{-t}(R\delta)$ which contradicts the assumption that a single tangency occurs. \qedd \\ There are three possible resolutions: first one of the two orbits, $\beta$ or $\gamma$ could undergo a pitchfork bifurcation creating a symmetry related pair of homoclinic orbits. For example, \Fig{fig:pitchfork} shows part of the homoclinic tangle at a parameter value where the type two homoclinic orbit with core sequence $(-\,-)$ pitchforks. As $k$ increases this results in the creation of a pair of type 2 orbits with cores $(-+)$ and $(+-)$, see \Fig{fig:pitch_plus}. Note that the new orbits are not symmetric, but that the reversal of $(-+)$ is $(+-)$, so they form a symmetric pair. \Epsfig{pitchfork.eps}{tbp} {Stable and unstable manifolds for the $\per{+}$ fixed point of the H\'enon map at $b=1$ and $k=3.09151$, where there is a cubic tangency of the manifolds at the $+^{\infty}-(-\,-)-+^{\infty}$ homoclinic orbit.} {fig:pitchfork}{3in} \Epsfig{pitchfork_plus.eps}{tbp} {Type $2$ homoclinic orbits of the H\'enon Map at $k=3.5$.} {fig:pitch_plus}{3in} The second possible bifurcation is a single-saddle node on the segment $\Wu(\beta,\gamma)$; this happens, for example, whenever a ``tip'' of an iterate of $S$ returns to $U$. This first occurs at type $3$, for the bifurcation $\bif{\sn}{(*-*)}$. We sketch a similar case, at type $4$, $\bif{\sn}{(*--*)}$, in \Fig{fig:asn} which occurs at $k\approx 3.982$. The third possible bifurcation is a pair of asymmetric saddle-node bifurcations. This first occurs for homoclinic orbits of type $4$. For example, the bifurcations $\bif{\sn}{\homo{+-+-},\homo{--+-}}$ and its time-reverse, $\bif{\sn}{\homo{-+-+},\homo{-+--}}$ occur at $k\approx 5.1886$. We sketch $U$ and $f^{-4}(S)$ at this bifurcation in \Fig{fig:asn}. This bifurcation also corresponds to the lower endpoint of an apparently hyperbolic parameter interval for the H\'enon map, as we discuss in the next section. Note that the antimonotonic bifurcations shown to exist in the area contracting case \cite{Yorke92} are exactly forbidden by this theorem. \Epsfig{asn.eps}{bt} {Sketch of two possible homoclinic saddle-node bifurcations of type four. A symmetric saddle-node creating $(*--*)$ occurs on $W^u((-++-),(++++))$ in (a). An asymmetric saddle-node creating $(*-+-)$ and $(-+-*)$ occurs with one point on $W^u((-^4),(+--+))$ in (b).} {fig:asn}{6in} A symmetric saddle-node followed by a pitchfork is a common bifurcation. For example, the parameter values, $k_{\rm sn}(t)$, at which the first type $t$ homoclinic orbit is created decrease monotonically with $t$. Thus at $k_{\rm sn}(t)$ the first type $t$ orbit is born and there are no homoclinic orbits with type less than $t$. For $t>1$, at $k_{\rm sn}(t-1)$ the segment $f^{t-1}(U)$ must intersect with $S$, so that $f^{t}(U)$ intersects with $f(S)$. In order for this to happen (when $b=1$), there is a pitchfork bifurcation for $k_{\rm pf}(t) \in [k_{\rm sn}(t), k_{\rm sn}(t-1)]$ of the type $t$ homoclinic orbit $\homoc{-+^{t-2}-}$ giving rise to the pair of orbits with symbol sequences \[ \ptcbif{\homo{-+^{t-2}-}}{\pf}{ \homo{-+^{t-1}} , \homo{+^{t-1}-}} \;. \] We see that the children of this bifurcation differ from their parent in a single symbol and they differ from each other in two symbols. \Tblref{tbl:pitch} lists the first few such homoclinic bifurcation values obtained by extrapolation of the first few members of the approximating orbit sequence. \begin{table}[tb] \centering { \begin{tabular}{c|c|l|c|l} \rule[-1.2mm]{0mm}{4mm} $t$ & Core & $k_{\rm sn}(t)$ & Pitchfork Children & $k_{\rm pf}(t)$ \\ \hline \rule[-1.2mm]{0mm}{4mm} 1 & $(*)$ & \:5.69931078670 & & \\ \rule[-1.2mm]{0mm}{4mm} 2 & $(**)$ & \:1.62777931098 & $(-+), (+-)$ & \:3.09150542113 \\ \rule[-1.2mm]{0mm}{4mm} 3 & $(*+*) $ & \:0.38555621701 & $(-+^2), (+^2-)$& \:0.71963023592 \\ \rule[-1.2mm]{0mm}{4mm} 4 & $(*+^2*)$ &-0.13347378530 & $(-+^3), (+^3-)$ &-0.04427324816 \\ \rule[-1.2mm]{0mm}{4mm} 5 & $(*+^3*)$ &-0.39678970175 & $(-+^4), (+^4-)$ &-0.36787481134 \\ \rule[-1.2mm]{0mm}{4mm} 6 & $(*+^4*)$ &-0.54918558488 & $(-+^5), (+^5-)$ &-0.53740149261 \\ \rule[-1.2mm]{0mm}{4mm} 7 & $(*+^5*)$ &-0.64623270965 & $(-+^6), (+^6-)$ &-0.64032496327 \\ \rule[-1.2mm]{0mm}{4mm} 8 & $(*+^6*)$ &-0.71262572399 & $(-+^7), (+^7-)$ &-0.70916824264\\ \rule[-1.2mm]{0mm}{4mm} 9 & $(*+^7*)$ &-0.76055766670 & $(-+^8), (+^8-)$ &-0.75830622014\\ \rule[-1.2mm]{0mm}{4mm} 10 & $(*+^8*)$ &-0.79659407362 & $(-+^9), (+^9-)$ &-0.79501732767\\ \end{tabular} } \caption{Pitchfork bifurcations from the first type $t$ orbits up to type 10.} \label{tbl:pitch} \end{table} The distance (in $k$) between the birth of the type $t$ orbit and its pitchfork bifurcation shrinks to zero as the type increases. \Sec{Intervals with no Bifurcations}{sec:entropy} Davis, MacKay and Sannami (DMS) \cite{Davis91} used the numerical method of Biham and Wenzel \cite{Biham89} to compute the periodic orbits for the area preserving H\'enon map. They showed that up to period $20$, there are intervals of parameter where there appear to be no orbits created or destroyed. They studied a particular parameter interval near the destruction of the horseshoe, and elucidated the symbolic dynamics of the corresponding homoclinic tangle. We will refer to this interval as the DMS gap. Though the method of Biham and Wenzel is guaranteed to work close enough to the AI limit \cite{Sterling98a}, it can fail \cite{Grassberger89}. We tested the DMS results using our continuation technique. The use of parallel computation allowed us to extend the original experiment by an order of magnitude in size so that we followed all orbits up to period $24$---recall from \Tblref{tbl:periods} that there are a total of $1{,}465{,}020$ possible orbits. We verify the DMS results and identify the symbol sequences of the orbits that form the boundaries of the DMS gap. \Epsfig{Plateaus.eps}{tb} {Number of periodic orbits of the H\'enon map up to period 24, for $b=1$. The maximal number is reached at $k\approx 5.69931078745$, when the horseshoe is formed. The endpoints of the largest gap studied by DMS are labeled by $L$ and $R$.} {fig:exitplateaus}{6in} In our experiment we follow all orbits up to period 24 and record the minimal parameter values at which they are destroyed. We then assume that each orbit exists only up to that value of $k$; this procedure is not entirely correct, because a few orbits loop back and forth in parameter under continuation. This is related to the vanishing of twist in the neighborhood of a period tripling bifurcation \cite{DMS99}. However, the number of low period orbits for which this happens is very small. In \Fig{fig:exitplateaus} we show the number of orbits that exist as a function of $k$, with the caveat that no value is plotted if the number of orbits does not change from the previously plotted point. This plot is equivalent to that of DMS, except that we leave gaps in the intervals where there are no bifurcations. At the anti-integrable limit the map exhibits a horseshoe so all of the periodic orbits are present. As we move away from the anti-integrable limit we see a decline in the number of periodic orbits as orbits collide and are destroyed. Flat intervals in \Fig{fig:exitplateaus} represent intervals of parameter where very few bifurcations occur. Gaps in the plot indicate intervals of parameter where there are {\em no} bifurcations. The creation of the first type $t$ homoclinic orbits gives rise to flat intervals. We observe that the left endpoint of each of the larger flat intervals for $k < 3$ corresponds to $k_{\rm sn}(t)$ for the saddle-node bifurcation of the first type $t$ homoclinic orbits; these are marked in \Fig{fig:exitplateaus} and in the enlargement, \Fig{fig:plateaus_zoom1}. Similarly, the parameter values $k_{\rm pf}(t)$ are also marked; note that these pitchfork bifurcations are located well beyond the right endpoints of the flat intervals. Each of the gaps in the flat intervals for $k < 3$ must eventually fill in if we go to high enough period because in this range of $k$ the area preserving H\'enon map has an elliptic fixed point. Recall that an $m/n$ bifurcation from the elliptic fixed point occurs at the parameter values $k_{m/n}$ given in \eq{Romega}, and these values are dense in the interval $-1 \le k \le 3$. Moreover, invariant circles bifurcate from the elliptic fixed point for each $k_{\omega}$ for sufficiently irrational $\omega$. The same argument can be used up to the end of the period doubling cascade of the fixed point at $k \approx 4.13616680392$, since each period doubling creates an elliptic orbit. There are a number of distinct gaps in \Fig{fig:exitplateaus}; the 3 larger gaps were studied by DMS, especially the largest one, near $k=5.5$ indicated by $L$ and $R$ in \Fig{fig:exitplateaus}. DMS conjecture that the dynamics in each gap is hyperbolic, and consequently there are no bifurcations in a gap. Our numerical evidence, which extends their study by an order of magnitude, supports this conjecture. Upon examining the orbits that limit on the endpoints of the gap up to period $24$, we can extrapolate and find that each of the five largest gaps is bounded by a homoclinic bifurcation, see \Tblref{tbl:gaps}. Thus we see that the gaps do not fill-in with orbits converging on the homoclinic bifurcations, but we cannot rule out that there are other, unrelated period orbits with period larger than $24$ that are created at parameter values in the middle of a gap. \begin{table}[tbp] \centering \begin{tabular}{ll|ll} Left Endpoint Core & $k_{L}$ & Right Endpoint Core & $k_{R}$ \\ \hline $(-+-*-+-) $ & 4.55931896797 & $(++-*-*-++)$ & 4.59567964802 \\ $(+^{3}-+-*)$ & 4.84317164217 & $(+^{3}-*-+-++-)$ & 4.86795762007 \\ $(-+-*) $ & 5.18851121215 & $(++-*-++)$ & 5.53765692812 \\ $(-++-*-++-)$ & 5.56490867348 & $(++-*-+^{3})$ & 5.60872105039 \\ $(-++-*)$ & 5.63190980280 & $(+^{3}-*-+^{3})$ & 5.67769222229 \\ \end{tabular} \caption{Homoclinic bifurcations bounding the gaps in \Fig{fig:exitplateaus}} \label{tbl:gaps} \end{table} We observe that there are two types of bifurcations bounding the gaps: symmetric and asymmetric saddle-node bifurcations. The asymmetric saddle-nodes result in the creation of two pairs of homoclinic orbits, the one listed in the table, and its time-reverse. Typically we observe that symmetric saddle node bifurcations in homoclinic orbits are followed by pitchfork bifurcations. In fact we observe that among all of the homoclinic orbits through type $11$, there are only $9$ special saddle-node bifurcations which are not followed by a pitchfork bifurcation. We believe that each of these special bifurcations corresponds to the endpoint of a gap. For example, the first type 1 bifurcation is not followed by a pitchfork, and it gives the left endpoint of the gap corresponding to the horseshoe. The four gap endpoints in \Tblref{tbl:gaps} that correspond to symmetric saddle nodes are each of this special type. The four remaining special pairs are each of type 11, and of these at least two bound gaps of widths $\Delta k \approx 6(10)^{-3}$. With our resolution it is not possible to clearly identify the final two as gap boundaries. In \Fig{fig:plateaus_zoom1} we show an enlargement of \Fig{fig:exitplateaus} but also include the data from the subshift, $\sigmaF$. In the upper right corner of \Fig{fig:plateaus_zoom1} we see the tail end of the exit time $2$ plateau. We also labeled the first large gap in $\sigmaF$ after the subshift is destroyed---this is the subshift analog of the DMS gap. \Epsfig{Plateaus_Zoom1.eps}{ht} {Enlargement: number of periodic orbits up to period 24 for the full shift and the subshift $\sigmaF$. The left and right ends of each plateau correspond to $k_{\rm sn}(t)$, and $k_{\rm pf}(t)$.} {fig:plateaus_zoom1}{6in} As in the DMS gap, the left and right boundaries, denoted $L$ and $R$, correspond to a pair of homoclinic saddle-node bifurcations with the core sequences \begin{eqnarray*} \label{subgap} \bif{\asn}{ \homo{*+-++-}, \homo{-++-+*} } & \mbox{ at\ } & k_{\rm sn}(L) \approx 1.533898312 \;, \\ \bif{\sn}{ \homo{+^3-++-+^3}, \homo{+^3-^4+^3} } & \mbox{ at\ } & k_{\rm sn}(R) \approx 1.583387630 \;. \end{eqnarray*} Note that the right endpoint of the gap corresponds to an orbit whose partner is not in the subshift! The curves for all orbits and for the subshift are remarkably similar and it appears that the growth of orbits in the subshift gives an accurate representation of the overall growth of orbits in the full shift for this range of parameters. This is especially remarkable given that when all the orbits exist, the subshift contains less than $1\%$ of the orbits in the full shift up to period 24. The figure shows that for small $k$, the number of orbits in the full shift is nearly a constant multiple of that in the subshift. Observing that the gaps are bounded by homoclinic orbits, we regenerated the orbit growth plot using {\it only} homoclinic orbits. As expected, the gap structure and overall shape of \Fig{fig:exitplateaus} is almost completely captured by the homoclinic orbits alone. \Sec{Conclusions}{sec:conc} Continuation from an anti-integrable limit is an effective technique for studying orbits providing that there are no isolated bubbles in the bifurcation diagram. In \cite{Sterling98a} we applied the anti-integrable theory to the H\'enon map to obtain a new proof of the well-known analytical bound of Devaney and Nitecki \cite{Devaney79}. In \Th{thrm:seq} we apply a similar argument to a restricted set of orbits to find an analytical bound for the existence of a subshift of finite type. We present both analytical bounds together with the optimal bounds generated numerically using our continuation method. We observe that the horseshoe is destroyed by a type one bifurcation that is homoclinic in the orientation preserving case, and heteroclinic otherwise. In either case we conjecture that this bifurcation is the first bifurcation among {\it all} orbits of the H\'enon map as we recede from the anti-integrable limit. With our continuation method, we are able to assign a ``global code'' to each orbit, by fixing the designation to that at the AI limit. In the H\'enon map, we demonstrate that this AI code is equivalent to the standard horseshoe code (when it exists), but it also gives a consistent way of assigning symbols to orbits {\it beyond} the destruction of the complete horseshoe. Remarkably, there appears to be a relationship between the AI codes for a number of systems including billiards, twist maps of the cylinder and the H\'enon map. We will explore this relationship in a forthcoming paper \cite{Sterling98b}. We relate the properties {\it transition time}, {\it type}, and {\it signature} of homoclinic orbits to properties of the core sequence. We also demonstrate that the ordering of the homoclinic orbits on the manifold segments $U$ and $S$ is the standard unimodal ordering. The notion of {\it double neighbors} and lemmas \ref{lem:doubleN} and \ref{lem:ttime} give a necessary condition for a pair of homoclinic orbits to bifurcate. Surprisingly, these also give a forcing relation that tells us which homoclinic bifurcations have to occur before other ones. Showing that homoclinic bifurcations can only take place between {\it double neighbors} with the same core length, \Lem{lem:neighbor} gives a symbolic criterion for a pair of homoclinic orbits to be neighbors. The ordering is certainly valid until the complete horseshoe is destroyed, which leads to the theorem that the first homoclinic bifurcation of the hyperbolic fixed point in the area preserving H\'enon map occurs between the pair of type one orbits. When $b= \pm 1$, the H\'enon map has a symmetry and we discuss the mechanism by which pitchfork and asymmetric saddle node bifurcations occur. The key ingredients to \Th{pitchfork} are the ordering on the manifolds and the existence of a reversor for the map. As a result, the scenario of a tip of the manifold just repeatedly piercing through the other manifold (which is most natural in the case without symmetry) is impossible. Among the possible alternatives are the occurrence of a pitchfork bifurcation or the creation of an asymmetric saddle-node pair, which are both not generic in the non-reversible case. With our continuation technique we compute numerical values for various bifurcations of the homoclinic orbits up to type eleven. We sketch the bifurcation diagram at type three, and then use a simple algorithm to construct the much more complex figures for higher core length. In contrast to our method for finding periodic orbits, the Biham and Wenzel method \cite{Biham90,Biham91} is known to fail in certain cases \cite{Grassberger89}, and can only be justified in the neighborhood of the AI limit \cite{Sterling98a}. Nevertheless, for the area preserving H\'enon map, we observe precisely the same number of orbits in the main DMS gap using our technique as was reported by DMS using the Biham-Wenzel method \cite{Davis91}. We extend the original experiment of DMS in two ways. First, we study an order of magnitude more orbits than the original experiment and yet the gaps originally reported by DMS persist. Second, we observe that homoclinic bifurcations are responsible for these gaps and we list the symbolic labels of the orbits that form the gap endpoints in Table \ref{tbl:gaps}. These gaps correspond to the creation and destruction of parameter intervals where the dynamics of the area preserving H\'enon map appears to be conjugate to a subshift of finite type. We find a similar gap structure for a particular subshift, $\sigmaF$, and list the symbolic labels for the homoclinic orbits that form the endpoints of the gap that is the analog of the DMS gap. The role of the special symmetric saddle-node bifurcations without an accompanying pitchfork bifurcation in this scenario remains to be elucidated. Many of our analytical results could be transferred from the area preserving case $b=1$ to the orientation reversing case $b=-1$. In this case there also exists a reversor, so that the theorem about the first bifurcation and about impossibility of pitchfork bifurcations could be generalized to this case. The main difference is that now we are not studying homoclinic orbits of a fixed point which is invariant under the reversor, but instead heteroclinic orbits connecting fixed points that are mapped into each other by the reversor. Correspondingly the action of the reversor on symbol sequences is not just reading them backwards, but it is reading backwards and flipping each symbol. Let us finally remark that we can extend the antimonotonicity result \cite{Yorke92} to the non-dissipative case $b=\pm 1$ because there are bifurcations that do occur in the ``wrong'' direction, i.e., orbits that are created when $k$ is decreased. This is described in detail in a separate paper \cite{DMS99} where it is related to the fact that in the neighborhood of the period tripling bifurcation the twist generically vanishes. \clearpage \bibliographystyle{unsrt}
\section{Introduction} Schur $Q$ functions first arose in the study of projective representations of $S_n$ \cite{sch-ur}. Since then they have appeared in variety of contexts including the representations of Lie superalgebras \cite{serg-eev} and cohomology classes dual to Schubert cycles in isotropic Grassmanians \cite{joze-fiak,prag-acz}. While studying the duality between skew Schur $P$ and $Q$ functions and their connection to the Schubert calculus of isotropic flag manifolds, we were led to their quasi-symmetric analogues: the \emph{peak functions} of Stembridge \cite {stem-bridge}. We show that \emph{the linear span of peak functions is a Hopf algebra} (Theorem~\ref{peak-coalg}). We also show that these peak functions are contained in the strictly larger set of \emph{shifted quasi-symmetric functions} (Theorem~\ref{kw-thetabh}) introduced by Billey and Haiman \cite{billey-haiman}. We remark that the quasi-symmetric functions here are not any apparent specialization of the quasi-symmetric $q$-analogues of Hivert \cite{Hivert}. {}From extensive calculations, we believe that the set of all shifted quasi-symmetric functions form a Hopf algebra, but at present we can only show that: \emph{The set of all shifted quasi-symmetric functions forms a graded coalgebra whose $n$th graded component has rank $\pi_n$, where $\pi_n$ is given by the recurrence} \[ \pi_n=\pi_{n-1}+\pi_{n-2} +\pi_{n-4},\] \emph{with initial conditions $\pi_1=1$, $\pi_2=1$, $\pi_3=2$, $\pi_4=4$.} We shall prove this result (Theorems~\ref{sqs-coalg} and~\ref{graded-rank}) and in addition shall establish some other properties of these functions. A composition $\alpha =[\alpha_1,\alpha_2, \ldots ,\alpha_k]$ of a positive integer $n$ is an ordered list of positive integers whose sum is $n$. We denote this by $\alpha \vDash n$. We call the integers $\alpha_i$ the \textit{components} of $\alpha$, and denote the number of components in $\alpha$ by $k(\alpha)$. There exists a natural one-to-one correspondence between compositions of $n$ and subsets of $[n-1]$. If $A=\{ a_1,a_2,\ldots ,a_{k-1}\}\subset[n-1]$, where $a_1<a_2<\ldots <a_{k-1}$, then $A$ corresponds to the composition, $\alpha =[a_1-a_0,a_2-a_1,\ldots ,a_k-a_{k- 1}]$, where $a_0=0$ and $a_k=n$. For ease of notation, we shall denote the set corresponding to a given composition $\alpha$ by $I(\alpha)$. For compositions $\alpha$ and $\beta$ we say that $\alpha$ is a \emph{refinement} of $\beta$ if $I(\beta)\subset I(\alpha)$, and denote this by $\alpha\preccurlyeq \beta$. For any composition $\alpha =[\alpha_1,\alpha_2,\ldots ,\alpha_k]$ we denote by $M_\alpha$ the \emph{monomial quasi-symmetric function} \cite{ges-sel} \[M_\alpha=\sum_{i_1<i_2<\ldots <i_k} x^{\alpha_1}_{i_1}\ldots x^{\alpha _k}_{i_k}.\] We define $M_0=1$, where $0$ denotes the unique empty composition of $0$. We denote by $F_\alpha$ the \emph{fundamental quasi-symmetric function} \cite{ges-sel} \[F_\alpha=\sum_{\alpha\preccurlyeq \beta} M_\beta .\] \begin{definition} For any subset $A\subset [n-1]$, let $A+1$ be the subset of $\{2, \dots , n\}$ formed from $A$ by adding $1$ to each element of $A$. Let $\alpha \vDash n$. Then we define \[\theta_\alpha = \sum_{\stackrel{\mbox{\scriptsize $\beta\vDash n$}}{I(\alpha)\subset I(\beta)\cup I(\beta)+1}}2^{k(\beta)}M_\beta .\] \label{st-em}\end{definition} This is the natural extension of the definition of peak functions given in~\cite{stem-bridge}. \begin{example} We shall often omit the brackets that surround the components of a composition. If $\alpha= 21$, then $I(\alpha)=\{ 2\}$, and $I(\alpha)+1=\{3\}$. Hence $$\theta _{21}=4M_{21}+4M_{12}+8M_{111}.$$ \end{example} Let $\Sigma^n$ be the $\mathbb{Z}$-module of quasi-symmetric functions spanned by $\{M_\alpha\}_{\alpha\vDash n}$ and let $\Sigma =\oplus _{n\geq 0}\Sigma^n$ be the graded $\mathbb{Z}$-algebra of quasi-symmetric functions. This is a Hopf algebra \cite{malv-reut} with coproduct given by \[\Delta(M_\alpha)= \sum_{\alpha =\beta \cdot\gamma}M_{\beta }\otimes M_{\gamma},\] where $\beta\cdot \gamma$ is the concatenation of compositions $\beta$ and $\gamma$. \begin{example} $\Delta (M_{32})= 1\otimes M_{32}+ M_3\otimes M_{2}+ M_{32} \otimes 1$. \end{example} We compute the coproduct of the functions $\theta_\alpha$. \begin{lemma} For any composition $\alpha\vDash n$ we have that \begin{eqnarray} \Delta (\theta_\alpha)&=& \sum \theta _{\epsilon\cdot a}\otimes\theta_{\phi(b\cdot\zeta)}\label{delta-eq}\end{eqnarray} where the sum is over all ways of writing $\alpha$ as $\varepsilon \cdot (a+b) \cdot \zeta$, that is, the concatenation of compositions $\varepsilon$ and $\zeta$, and a component of $\alpha$ written as the sum of numbers $a,b\geq0$. Also $\phi(b\cdot\zeta)= [1+\zeta_1,\zeta_2,\ldots]$ if $b=1$ and $b\cdot\zeta$ otherwise. \label{mult-coprod} \end{lemma} We shall use this result to show that certain subsets of functions $\theta_\alpha$ span coalgebras (Theorems~\ref{peak-coalg} and~\ref{sqs-coalg}). \begin{proof} Definition~\ref{st-em} is equivalent to \[ \theta_\alpha =\sum_{\stackrel{\mbox{\scriptsize $\beta\vDash n$}}{\mbox{\scriptsize $\beta^\ast \preccurlyeq \alpha$}}} 2^{k(\beta)} M_\beta , \] where $\beta^\ast$ is the refinement of $\beta$ obtained by replacing all components $\beta_i >1$, for $i>1$, by $[1,\beta_i-1]$. Thus the LHS of equation (\ref{delta-eq}) is equal to \begin{eqnarray} \sum_{\stackrel{ \stackrel{\mbox{\scriptsize$\beta\vDash n$}} {\mbox{\scriptsize$\beta^\ast\preccurlyeq\alpha$}}} {\mbox{\scriptsize$\beta=\gamma\cdot\delta$}}} 2^{k(\beta)} M_\gamma \otimes M_\delta &=&\sum_{\stackrel{\mbox{\scriptsize$\gamma\cdot\delta\vDash n$}}{(\gamma\cdot\delta)^\ast\preccurlyeq\alpha}} 2^{k(\gamma)}M_\gamma\otimes 2^{k(\delta)}M_\delta. \label{2delta-eq} \end{eqnarray} Let $2^{k(\gamma)}M_\gamma\otimes 2^{k(\delta)}M_\delta$ be a term of this sum, with $\gamma\vDash m$. This term can only appear in one summand on the RHS of equation (\ref{delta-eq}), namely $\theta_{\varepsilon\cdot a}\otimes \theta_{\phi(b\cdot\zeta)}$ with $\varepsilon\cdot a\vDash m$. To show that it does indeed appear, we need to prove that $\gamma^\ast \preccurlyeq \varepsilon\cdot a$ and $\delta^\ast \preccurlyeq \phi(b\cdot\zeta)$. Let $\delta^{\ast\ast}$ be the refinement of $\delta^\ast$ obtained by replacing the part $\delta_1$ by $[1,\delta_1-1]$ if $\delta_1>1$. We have that \[\gamma^\ast\cdot \delta^{\ast\ast}=(\gamma\cdot\delta)^\ast\preccurlyeq \varepsilon\cdot(a+b)\cdot\zeta ,\] which implies that $\gamma^\ast \preccurlyeq \varepsilon\cdot a$, and $\delta^{\ast\ast}\preccurlyeq b\cdot\zeta\preccurlyeq \phi(b\cdot \zeta)$. If $\delta_1=1$ then $\delta^\ast =\delta^{\ast\ast}\preccurlyeq\phi (b\cdot \zeta)$. However, if $\delta _1>1$ then there are two possible cases: either $\delta_1\leq b$, or $b=1$ and $\delta_1-1\leq\zeta_1$. In the former case $\delta^\ast \preccurlyeq b\cdot\zeta =\phi (b\cdot \zeta)$, while in the latter, $\delta _1\preccurlyeq 1+\zeta_1$, whence $\delta^\ast \preccurlyeq [1+\zeta_1,\zeta_2,\ldots ]= \phi ( b\cdot\zeta)$. Conversely, let $ 2^{k(\gamma)}M_\gamma\otimes 2^{k(\delta)} M_\delta$ be a term belonging to a tensor $\theta_{\varepsilon\cdot a}\otimes \theta_{\phi(b\cdot \zeta)}$ on the RHS of equation (\ref{delta-eq}). To show that it appears in equation (\ref{2delta-eq}) we must prove that $(\gamma\cdot \delta )^\ast\preccurlyeq \varepsilon\cdot (a+b)\cdot \zeta$. We have that $\gamma^\ast\preccurlyeq \varepsilon\cdot a$ and $\delta^\ast\preccurlyeq \phi(b\cdot \zeta)$, which imply that \[(\gamma\cdot \delta)^\ast= \gamma^\ast\cdot\delta^{\ast\ast}\preccurlyeq \gamma^\ast\cdot\delta^\ast \preccurlyeq \varepsilon\cdot a\cdot \phi(b\cdot \zeta).\] If $b>1$ then $$(\gamma\cdot\delta)^\ast\preccurlyeq \varepsilon\cdot a\cdot \phi(b\cdot \zeta)=\varepsilon\cdot a\cdot b\cdot \zeta\preccurlyeq \varepsilon\cdot (a+b)\cdot\zeta.$$ If $b=1$ then $\delta^\ast \preccurlyeq \phi(b\cdot\zeta)=[1+\zeta_1,\zeta_2,\ldots ]$ implies that \[\delta^{\ast\ast}=[1, \ldots ]\preccurlyeq [1,\zeta_1,\ldots ]=b\cdot \zeta .\] Therefore, \[(\gamma\cdot\delta)^\ast=\gamma^\ast\cdot\delta^{\ast\ast}\preccurlyeq \varepsilon\cdot a\cdot b\cdot\zeta\preccurlyeq \varepsilon\cdot (a+b)\cdot \zeta\] as desired. \end{proof} \section{The peak Hopf algebra} \begin{definition} For any composition $\alpha =[\alpha_1,\alpha_2,\ldots ,\alpha_k]$ we say that $\theta_\alpha$ is a \emph{peak function} if $\alpha_i = 1\Rightarrow i=k$.\end{definition} Observe that if $\theta_\alpha$ is a peak function and $\alpha\vDash n$, then $I(\alpha)\subset \{2,\ldots ,n-1\}$ such that no two $i$ in $I(\alpha)$ are consecutive. Let $\Pi^n$ be the $\mathbb{Z}$-module spanned by all peak functions $\theta _\alpha$, $\alpha\vDash n$, and let $\Pi=\oplus_{n\geq 0}\Pi^n$. This was studied by Stembridge \cite{stem-bridge} who showed that the peak functions are F-positive, are closed under product, and form a basis for $\Pi$, and so the rank of $\Pi^n$ is the $n$th Fibonacci number. In addition we also know the following about the \emph{algebra of peaks}, $\Pi$. \begin{theorem} $\Pi$ is closed under coproduct. \label{peak-coalg}\end{theorem} \begin{proof} If all components of a composition $\alpha$, except perhaps the last, are greater than $1$, then the same is true for all compositions $\varepsilon\cdot a$ and $\phi(b\cdot\zeta)$ appearing in the RHS of equation (\ref{delta-eq}). \end{proof} Let $\Theta$ be the $\mathbb{Z}$-linear map from $\Sigma $ to $\Pi $ defined by $\Theta(F_\alpha)=\theta_{\Lambda(\alpha)}$, where $\Lambda(\alpha)$ is the composition formed from $\alpha =[\alpha_1,\alpha_2,\ldots ,\alpha _k]$ by adding together adjacent components $\alpha_i ,\alpha_{i+1},\ldots , \alpha_{i+j}$ where $\alpha_{i+l}=1$ for $l=0,\ldots , j-1$, and either $\alpha_{i+j}\neq 1$, or $i+j=k$. \begin{example} If $\alpha=31125111$ then $\Lambda (\alpha)=3453$. \end{example} Stembridge \cite{stem-bridge} showed that $\Theta:\Sigma \rightarrow \Pi$ is a graded surjective ring homomorphism, and was an analogue of the retraction from the algebra of symmetric functions to Schur $Q$ functions. It is clear from our proof above that this morphism is in fact a Hopf homomorphism. We can describe the kernel of $\Theta$ as follows. \begin{lemma} The non-zero differences $F_\alpha-F_{\Lambda (\alpha)}$ form a basis of the kernel of $\Theta$. \end{lemma} \begin{proof} Each difference $F_\alpha -F_{\Lambda(\alpha)}$ is in the kernel of $\Theta$ as $\Theta(F_\alpha -F_{\Lambda(\alpha)})=0$ since $\Lambda(\Lambda(\alpha))=\Lambda(\alpha)$. In addition, the non-zero differences are linearly independent as they have different leading terms. Letting $f_n$ denote the $n$th Fibonacci number, there are $2^{n-1}-f_n$ such differences, and since \begin{eqnarray*} \dim\ker\Theta &=& \dim \Sigma^n - \dim \Pi^n\\ &=&2^{n-1} -f_n, \end{eqnarray*} our result follows. \end{proof} \section{The coalgebra of shifted quasi-symmetric functions} \begin{definition} For any composition $\alpha =[\alpha_1,\alpha_2,\ldots ,\alpha_k]\vDash n$ we say that $\theta_\alpha$ is a \emph{shifted quasi-symmetric function} (sqs-function) if $n\le 1$ or $\alpha_1> 1$.\end{definition} Observe that if $\theta_\alpha$ is an sqs-function and $\alpha\vDash n$, then $I(\alpha)\subset \{2,\ldots ,n-1\}$. For integers $n\geq 0$, let $\Xi^n$ be the $\mathbb{Z}$-module spanned by all sqs-functions $\theta_\alpha$, $\alpha\vDash n$, and let $\Xi=\oplus_{n\geq 0}\Xi^n$. \begin{theorem} $\Xi$ is closed under coproduct. \label{sqs-coalg} \end{theorem} \begin{proof} If the first component of a composition $\alpha$ is greater than $1$, then the same is true for all compositions $\varepsilon\cdot a$ and $\phi(b\cdot\zeta)$ appearing in the RHS of equation (\ref{delta-eq}). \end{proof} Unlike peak functions \cite{stem-bridge}, sqs-functions are not $F$-positive since $$\theta_{211}=F_{22}+ F_{112}+ 2F_{121}+ F_{211}- F_{1111}.$$ \begin{definition} For any composition, $\alpha\vDash n$, we define the complement $\alpha^c$ of $\alpha$ to be the composition for which $I(\alpha^c) = (I(\alpha))^c$, the set complement of $I(\alpha)$ in $[n-1]$. We define the graph $G(\alpha)$ of $\alpha$ to be the graph obtained from $$ \epsfxsize=1.8in\epsfbox{graphn.eps} $$ by removing the edge $(i,i+1)$ if and only if $i\in I(\alpha)$. \end{definition} Observe that $G(\alpha^c)$ contains the edge $(i,i+1)$ if and only if this edge is not contained in $G(\alpha)$. These graphs will be used later to simplify the proof of Theorem~\ref{kw-thetabh}. Let a \emph{word} of length $n$ be any $n$-tuple, $w_1w_2\ldots w_n$, and let a \emph{binary word} of length $n$ be a word $w_1w_2\ldots w_n$ wuch that $w_i\in \{ 0,1\}$ for all $i$. For $2\leq i\leq n-1$, let us denote by $\ensuremath{3^{(i)}} $ the composition $ [1^ {i-2}, 3,1^{n-i-1}]$ of $n $. For some subset $S\subset \{2, \ldots ,n-1\}$, let us denote by $\bigwedge_{i\in S}\ensuremath{3^{(i)}}$ the composition of $n$ for which $G(\bigwedge_{i\in S}\ensuremath{3^{(i)}})$ has an edge between vertices $i$ and $i+1$ if and only if an edge exists between vertices $i$ and $i+1$ in $G(\ensuremath{3^{(i)}})$ for some $i\in S$. \begin{example} Let $S=\{ 2,3\}\subset [3]$. Then $G(3^{(2)})$ is $$ \epsfxsize=1.55in\epsfbox{graph2.eps} $$ and $G(3^{(3)})$ is $$ \epsfxsize=1.55in\epsfbox{graph3.eps} $$ hence $G(\bigwedge_{i\in S}\ensuremath{3^{(i)}})$ is $$ \epsfxsize=1.55in\epsfbox{graph4.eps} $$ so $\bigwedge_{i\in S}\ensuremath{3^{(i)}}$ is the composition $4$. \end{example} \begin{definition}\cite{billey-haiman} Let $\alpha$ be a composition of $n$. Let ${\mathcal A}(I(\alpha))$ denote the set of all sequences $j_1\leq j_2\leq\ldots \leq j_n$ in $\mathbb{N}$ such that we do not have $j_{i-1}=j_i=j_{i+1}$ for any $i\in I(\alpha)$. The \emph{shifted quasi-symmetric function} $\ensuremath{\theta^{BH}}_\alpha$ is given by \[\ensuremath{\theta^{BH}}_\alpha = \sum_{\stackrel{\mbox{\scriptsize$J=(j_1,\ldots, j_n)$}}{ \stackrel{\mbox{\scriptsize$j_1\leq\ldots\leq j_n$}}{ J\in{\mathcal A}(I(\alpha))}}} 2^{| j|}x_{j_1}\ldots x_{j_n},\]where $|j|$ denotes the number of distinct values $j_i$ in $J$.\label{bi-ha} \end{definition} \begin{theorem} For any sqs-function $\theta_\alpha$ we have that $\theta_\alpha=\ensuremath{\theta^{BH}} _\alpha$.\label{kw-thetabh} \end{theorem} \begin{proof} For each $i\in \ensuremath{{I(\alpha)}}\subset[n-1]$, $j_{i-1}=j_i=j_{i+1}$ is forbidden in any monomial \[ x_{j_1}x_{j_2}\ldots x_{j_i}\ldots x_{j_n}\] appearing as a summand of the function $\ensuremath{\theta^{BH}}_\alpha$. This is equivalent to saying that $M_\beta$ is a summand of $\ensuremath{\theta^{BH}}_\alpha$ if and only if $G(\ensuremath{3^{(i)}})\not\subset G(\beta)$ for all $i\in \ensuremath{{I(\alpha)}}$. Therefore at least one of $i-1$ or $i$ must be the largest label of a vertex in a connected component in $G(\beta)$. Now when going from compositions of $n$ to subsets of $[n-1]$ we can do so using our graphs, $G$. All we have to do is list the label of the vertex that is the largest in each connect component, not listing $n$. We call these vertices the \emph{end-points}. We are now in a position to prove the equivalence of Definitions ~\ref{st-em} and ~\ref{bi-ha} for sqs-functions. The powers of 2 agree so we need only show that the indices of summation do too. To see this, take any sqs-function $\theta_\alpha$ and let $i\in\ensuremath{{I(\alpha)}}$. Then $M_\beta$ is a summand in $\ensuremath{\theta^{BH}}_\alpha$ if at least one of $i-1$ or $i$ is an end-point in $G(\beta)$. Therefore $i$ or $i-1$ belongs to $I(\beta)$, and $M_\beta$ is a summand of $\theta_\alpha$. Conversely, if $M_\beta $ is a summand of $\theta _\alpha$, then this implies that for each $i\in \ensuremath{{I(\alpha)}}$, we have that $i-1$ or $i$ belongs to $I(\beta)$, so one of $i-1$ or $i$ is an end-point in $G(\beta)$, so $M_\beta$ is a summand of $\ensuremath{\theta^{BH}}_\alpha$. \end{proof} \section{A basis for $\Xi$} \begin{definition} Let $\theta_\alpha$ be an sqs-function and $\alpha \vDash n$. We define an internal peak $i\in \ensuremath{{I(\alpha)}}$ such that $i- 1,i+1\not\in \ensuremath{{I(\alpha)}}$, and $i\in\{ 3,\ldots n-2\}$. \end{definition} \begin{remark} Observe that the occurrence of an internal peak in the $i$th position in $I(\alpha)=\{ w_1,w_2,\ldots\}$, where $w_1<w_2<\ldots$, is equivalent to having two components of $\alpha$, say $\alpha_i,\alpha_{i+1}$ such that $\alpha_{i+1}\geq 2$, and $\alpha_i\geq 2$ if $i\neq1$, or $\alpha_i\geq 3 $ if $i=1$. \end{remark} We can now describe the basis of $\Xi$ as follows. \begin{theorem} The coalgebra $\Xi$ has a basis consisting of all sqs-functions $\theta_\alpha$ where $\ensuremath{{I(\alpha)}}$ contains no internal peak.\label{xi-basis} \end{theorem} We sketch the proof of Theorem 4.2 later. \begin{theorem} The rank of $\Xi^n$ is given by the recurrence \[ \pi_n=\pi_{n-1}+\pi_{n-2} +\pi_{n-4},\] with initial conditions $\pi_1=1$, $\pi_2=1$, $\pi_3=2$, $\pi_4=4$.\label{graded-rank} \end{theorem} This recurrence was suggested by a superseeker query \cite{sloan-seeker}. \begin{proof} By direct calculation we obtain that $\pi_1=1$, $\pi_2=1$, $\pi_3=2$, and $\pi_4=4$. To obtain our recurrence, we observe that for each sqs-function, $\theta_\alpha$ where $\alpha\vDash n$, we can encode $I(\alpha)$ as a binary word of length $n-2$, by placing a $1$ in position $i-1$ if $i$ is contained in $I(\alpha)$, and $0$ otherwise. By this one-to-one correspondence we see that $I(\alpha)$ contains no internal peak if its corresponding binary word does not contain $010$ as a subword. We therefore count binary words of length $n$ that avoid the subword $010$. Appending either $1$ or $0$ to such a binary word of length $n-1$ gives one of length $n$, provided that we have not created the subword $010$ in the last three positions. Let $a_n$, $b_n$, $c_n$, and $d_n$ enumerate those binary words of length $n-2$ that avoid the subword $010$ and end in, respectively $00$, $01$, $10$, and $11$. We then obtain the following 4 simultaneous recursions. \[a_n=a_{n-1} +c_{n-1}, \ b_n=a_{n-1} +c_{n-1}, \ c_n=d_{n-1}, \ d_n=b_{n-1} +d_{n-1}. \] Clearly the number of $I(\alpha)$s in $[n-1]$ with no internal peaks is given by \[ \pi_n=a_n+b_n+c_n+d_n,\] However by substituting in our recurrences we obtain \begin{eqnarray*} \pi_n&=& a_n+b_n+c_n+d_n\\ &=& 2a_{n-1}+b_{n-1}+2c_{n-1}+2d_{n-1}\\ &=& \pi_{n-1}+a_{n-1}+c_{n-1}+d_{n-1}\\ &=& \pi_{n-1}+a_{n-2}+b_{n-2}+c_{n-2}+2d_{n-2}\\ &=& \pi_{n-1}+\pi_{n-2} +d_{n-2}\\ &=& \pi_{n-1}+\pi_{n-2} +b_{n-3} +d_{n-3}\\ &=& \pi_{n-1}+\pi_{n-2} +a_{n-4}+b_{n-4} +c_{n-4}+d_{n-4}\\ &=& \pi_{n-1}+\pi_{n-2} +\pi_{n-4}. \\ \end{eqnarray*} \end{proof} We say that $M_\beta$ is a maximal term of $\theta_\alpha$ if for any $\gamma$ higher in the partial order of compositions $M_\gamma$ is not a summand of $\theta_\alpha$. The following lemma is stated without proof. \begin{lemma} Let $\theta_\alpha$ be an sqs-function. Consider the collection $S$ of all possible sets derived from $I(\alpha)$ by adding either $i-1$ or $i+1$ to $\ensuremath{{I(\alpha)}}$ for all internal peaks $i\in \ensuremath{{I(\alpha)}}$. If $M_\beta$ is a maximal term of $\theta_\alpha$, then $\beta$ is derived from \[\bigwedge_{\stackrel{\mbox{\scriptsize$i\in (\ensuremath{I(\tilde{\alpha})} )^c$}}{\ensuremath{I(\tilde{\alpha})}\in S}} \ensuremath{3^{(i)}} \] by adding adjacent components equal to 1 together to give a component equal to 2 as often as possible. \label{big-term} \end{lemma} \begin{lemma} Let $\theta_\alpha$ be an sqs-function, and let $ \ensuremath{{I(\alpha)}}$ have an internal peak in the $j$th position, then we have the following linear relation \begin{eqnarray*}\theta_\alpha &=&\theta_{[\alpha_1,\ldots ,\alpha_j-1,1,\alpha_{j+1},\ldots ,\alpha_k]}+ \theta_{[\alpha_1,\ldots ,\alpha_j,1,\alpha_{j+1}-1,\ldots ,\alpha_k]}\\ &&- \theta_{[\alpha_1,\ldots ,\alpha_j-1,1,1,\alpha_{j+1}-1,\ldots ,\alpha _k]}.\end{eqnarray*}\label{rel-ation} \end{lemma} \begin{proof} By Definition~\ref{bi-ha} we have that the leading terms of $\theta _\alpha$ determine the other summands that belong to $\theta_\alpha$. Hence by Lemma~\ref{big-term} it follows that the summands of $\theta _\alpha$ will be the union of the summands of $ \theta_{[\alpha_1,\ldots ,\alpha_j-1,1,\alpha_{j+1},\ldots ,\alpha_k]}$ and $\theta_{[\alpha _1,\ldots ,\alpha_j,1,\alpha_{j+1}-1,\ldots ,\alpha_k]}$. However, those summands that appear in both will be duplicated. By definition these will be the summands of $\theta_{[\alpha_1,\ldots ,\alpha_j-1,1,1,\alpha _{j+1}-1,\ldots ,\alpha_k]}$, and the result follows. \end{proof} \noindent\textit{Sketch of proof of Theorem~\ref {xi-basis}.} {}From our relation in Lemma~\ref{rel-ation}, it follows that any $\theta_\alpha$ can be rewritten as a linear combination of functions $\theta_{\tilde{\alpha}}$, where $I(\tilde{\alpha})$ contains no internal peaks. In addition, by Lemma~\ref{big-term} and definition~\ref{bi-ha} we have that the set of all sqs-functions $\theta_\alpha$ where $I(\alpha)$ contains no internal peaks is linearly independent and thus form a basis for $\Xi$.\qed
\section{Introduction} Except for the $^{12}$CO(1--0) line, few detailed maps of the molecular gas distribution in galaxies have been published so far. $^{13}$CO(1--0) has been rarely observed in galaxies, and $^{12}$CO(2--1) even more rarely. The weakness of the $^{13}$CO(1--0) line in galaxies and the technical problems of observing at the 230\,GHz frequency of $^{12}$CO(2--1) account for this rarity of observations. However, for a better knowledge of the interstellar medium in galaxies, more tracers than $^{12}$CO(1--0) must be observed. A crucial issue that can be studied through these observations is the validity of the conversion factor $M(\rm H_2)/\mathit L_{\rm CO}\simeq 4.8$\,\mo\,(K\,$\kms$\,pc$^2$)$^{-1}$ that is valid for self-gravitating molecular clouds (see Sa\-ge \& Isbell 1991). $^{13}$CO(1--0) has been observed with interferometers in only a few galaxies. In IC\,342, a galaxy similar to the Milky Way, Wright et al. (1993) showed a variation of the intensity ratio $R_{12/13}=^{12}$CO(1--0)/$^{13}$CO(1--0). Those authors interpret $^{13}$CO peaks as molecular clouds while $^{12}$CO would trace a more diffuse medium. The variations of $R_{12/13}$ can be interpreted in different ways (Sage \& Isbell 1991, Sakamoto et al. 1997), including variation of gas density, kinetic temperature, and relative abundance of isotopes. Downes et al. (1992) explained these ratio variations by filling factors of clouds varying across the central part of IC\,342. Up to now $^{12}$CO(2--1) has been observed mainly in the GMCs of our Galaxy, where the ratio $R_{21/10}=^{12}$CO(2--1)/$^{12}$CO(1--0) has a typical value of 1.0 (Plam\-beck \& Wil\-liams 1979), indicating gas at low kinetic temperature ($\sim 10$\,K). Radford et al. (1991) observed a low ratio $R_{21/10}\simeq 0.6-0.75$ in infrared-luminous galaxies, indicating the presence of subthermally excited CO. Braine et al. (1993) surveyed of 81 galaxies in the CO transitions, and found a average ratio $R_{21/10}\simeq 0.89$, indicating cold, optically thick gas. Similarly in the central region of IC\,342, Eckart et al. (1990) found a ratio $R_{21/10} \approx 1.0$ everywhere, with a slightly higher value in the center, possibly indicating an increase in the gas kinetic temperature toward the center. A ratio $R_{21/10} \approx 1.4$ was found close to a CO arm, indicating warm gas heated by star formation. The starburst galaxy M82 shows a different behaviour, with an unusually high ratio $R_{21/10} \approx 2.5$ (Knapp et al. 1980, Loiseau et al. 1990), indicating hot gas ($>40\,$K) heated by star formation. We have observed in the $^{13}$CO(1--0) and $^{12}$CO(1--0) lines the central region ($\simeq 5\,$kpc) of the barred spiral galaxy {NGC$\,$1530}. This galaxy contains large amount of molecular gas in its central kiloparsec, due to the accreting action of its bar (cf models by Athanassoula 1992, Friedli \& Benz 1993, Piner, Stone, \& Teuben 1995). The bar has driven a high fraction ($\simeq 25\%$) of the total gas of the galaxy into the center (Downes et al. 1996, DRSR hereafter). CO(1--0) has been extensively studied in this galaxy (Regan et al. 1995, DRSR, Reynaud \& Downes 1997 ({\rm RD}\ 97), Reynaud \& Downes 1998). {\rm RD}\ 97 also mapped HCN(1--0), showing that the dense gas as traced by HCN is mainly concentrated in a nuclear ring or unresolved spiral at galactic radius $\leq 1$\,kpc. This concentration of dense gas is connected with the presence of an inner Lindblad resonance at radius $\simeq 1.2$\,kpc. The ionized gas was mapped in H$\alpha$ by Regan et al. (1996). Greve et al. (1999) compared the distribution and kinematics of ionized and molecular gas in the bar of {NGC$\,$1530}. \section{Observations, Data reduction} The observations were made with the IRAM interferometer on Plateau de Bure, France, with four 15m antennas (Guilloteau et al. 1992) between November 1995 and April 1996. Each antenna of the interferometer was equip\-ped with a dual channel receiver tuned to frequencies 109.3\,GHz and 228.7\,GHz, corresponding to $^{13}$CO(1--0) and $^{12}$CO(2--1). The tracking center of the interferometer was $\ra[04 23 27.32]$, $\dec[+75 17 45.0]$ (J2000), close to the nucleus of the galaxy. The recession velocity of the galaxy is $\simeq 2470\kms$. See DRSR for a table with the astrophysical parameters of the galaxy. The interferometer was used in 3 configurations, giving baselines between 20m and 180m. After Fourier transform of the visibilities and restoration by the CLEAN algorithm, we obtained channel maps of both transitions with a $20\kms$ velocity width. The observational parameters are summarised in Table~\ref{table:obs}, including the clean beam parameters and the final noise in channel maps. \begin{table} \caption{IRAM interferometer parameters.} \begin{tabular}{c|c c} \hline Parameter & $^{13}$CO(1--0) & $^{12}$CO(2--1) \\ \hline Frequency (GHz) & 109.3 & 228.7 \\ Calibrator flux (Jy) $^a$ & 1.1 & 0.5 \\ Beam (FWHM) & $3''.4 \x 2''.9$ & $1''.8\x 1''.4$ \\ Beam (p.a.) & $-76\deg$ & $-122\deg$ \\ $T_{\rm b}/S$ (K/Jy) & 10.4 & 9.0 \\ r.m.s. noise (mJy\,beam$^{-1}$) $^b$ & 1.5 & 5.0 \\\hline \multicolumn{3}{l}{{\scriptsize \it a} Source=0224+671}\\ \multicolumn{3}{l}{{\scriptsize \it b} Noise in $20\kms$ channels}\\ \end{tabular} \label{table:obs} \end{table} \section{Observational Results} \subsection{Molecular maps} Figure~\ref{fig:molecules} shows the integrated $^{13}$CO(1--0) and $^{12}$CO(2--1) of the inner $30''$ of {NGC$\,$1530}. In the same figure are also displayed for comparison the integrated $^{12}$CO(1--0) and HCN(1--0) maps from {\rm RD}\ 97. The integration ranges are respectively $540\kms$ ($^{12}$CO(1--0)), $520\kms$ ($^{12}$CO(2--1)), $360\kms$ ($^{13}$CO(1--0)), and $180\kms$ (HCN(1--0)). Each map is corrected for the primary beam attenuation. The primary beams (FWHM) are respectively $43''$ ($^{12}$CO(1--0)), $21''$ ($^{12}$CO(2--1)), $45''$ ($^{13}$CO(1--0)), and $56''$ (HCN(1--0)). These maps are centered on the dynamical center of the galaxy, which has coordinates $\ra[04 23 26.7]$, $\dec[75 17 44.0]$ (J2000) (from RD 97), i.e. $2.5''$ from the tracking center of the interferometer. \begin{figure*}[p] \centerline{\resizebox{1.0\hsize}{!}{\includegraphics[angle=270]{8337.f1}}} \vspace*{1cm} \centerline{\resizebox{0.3\hsize}{!}{\includegraphics[width=6cm,angle=270]{8337.f1bis}}} \caption{False color maps of four velocity integrated transitions. {\em a)} upper left: $^{12}$CO(1--0). Labels indicate the regions described in section~4.1. A sketch of the arcs and the nuclear feature (ring or unresolved spiral arms displayed as an ellipse) is also displayed (full black line). {\em b)} upper right: $^{12}$CO(2--1). {\em c)} lower left: $^{13}$CO(1--0). {\em d)} lower right: HCN(1--0). For each map the {\bf X} sign indicates the position of the dynamical center, and the clean beam is indicated at lower left. Each map is corrected for the primary beam attenuation (see text). The primary beams (FWHM) are shown as red circles. The color scale is indicated at the right of the diagram. The minimum displayed flux are 0 for all transitions, except for $^{12}$CO(2--1) where it is 3\,Jy\,beam$^{-1}\kms$. The maxima correspond to the value 1 on the color scale, and are 10.0\,Jy\,beam$^{-1}\kms$ ($^{12}$CO(1--0)), 25.0 ($^{12}$CO(2--1)), 2.7 ($^{13}$CO(1--0)) and 1.7 (HCN(1--0)) respectively. The bottom panel shows an optical image of the bar and a part of the spiral arms of NGC\,1530 (Optical image NOAO). The red square indicates the region shown in the four integrated transitions.} \label{fig:molecules} \end{figure*} The $^{12}$CO(2--1) map is truncated beyond a diameter of $42''$. Structures visible at the truncation limit are probably real molecular clouds deformed by the high noise level, since the noise is amplified by the primary beam correction. The two transitions of $^{12}$CO give similar maps, with two arcs, and inside them, a central structure which is a ring or unresolved nuclear spiral arms. These two maps have similar resolutions, $1''.8$ and $1''.6$ for $^{12}$CO(1--0) and $^{12}$CO(2--1) respectively. The $^{13}$CO(1--0) map is grossly similar to the $^{12}$CO maps, with a beam twice as large ($3''.1$). However the arcs seem dimmer in $^{13}$CO than in $^{12}$CO. There is a real difference between the brightness of the arcs and the brightness of the nuclear feature, a difference which had already been detected in HCN (see Fig.~\ref{fig:molecules} and Fig.~3 of {\rm RD}\ 97). The difference is confirmed on the ratio map obtained by smoothing $^{12}$CO(1--0) to the resolution of $^{13}$CO(1--0) (Fig.~\ref{fig:12_13}). Figure~\ref{fig:13co10_channel} shows the $20\kms$ channel maps of the $^{13}$CO(1--0) emission. Figure~\ref{fig:12co21_channel} shows the channel maps of the $^{12}$CO(2--1) emission. These maps are not corrected for the primary beam attenuation. The kinematic pattern shown by these maps is the same as that found by {\rm RD}\ 97 in the $^{12}$CO(1--0) transition. That is, the kinematics of the gas in the arcs shows large ($100\kms$) infall motions (due to the $x_1$ orbits along the bar) and in the central feature shows mainly circular rotation or weakly elliptical orbits (the $x_2$ orbits normal to the bar). Figure~\ref{fig:posvel} shows position-velocity diagrams in the CO(2--1) (left panel) and the CO(1--0) (right panel) transitions. These diagrams are cuts in the data cube, along the line of nodes passing through the dynamical center. The circular component is therefore the only component of the velocity field detected on these diagrams. The maximal radius of the emission is 1.4\,kpc. The diagrams are very similar in the two transitions of CO, with a steep rising part in the central $3''$ region and a flattening of the rotation curve at the crossing of the nuclear feature (incomplete ring or spiral within a $6''$ diameter of the nucleus). Outside this region, the rotation curve is steep (see Fig.~\ref{fig:posvel} at radii of $8''$). \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[angle=270]{8337.f2}} \caption{$^{13}$CO(1--0) maps of the central $25''$ of {NGC$\,$1530}\ in $20\kms$ wide channels. Radial velocities ($\kms$, upper left of each box) are relative to $2450\kms$. The contour intervals are $-6$, $-3$, 3, 6, 9, 12, 21, 30, 39, 48\,mJy\,beam$^{-1}$ ($\sigma =1.5\,$mJy\,beam$^{-1}$). The cross indicates the position of the tracking center of the interferometer ($\alpha=\ra[4 23 27.3]$, $\delta=\dec[75 17 45.0]$; J2000). The $3''.4\x 2''.9$ clean beam is shown in the lower right box.} \label{fig:13co10_channel} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[angle=270]{8337.f3}} \caption{$^{12}$CO(2--1) maps of the central $25''$ of {NGC$\,$1530}\ in $20\kms$ wide channels. Radial velocities ($\kms$, upper left of each box) are relative to $2429\kms$. The contour intervals are $-60$, $-30$, 30, 60, 90, 120, 180, 240, 300\,mJy\,beam$^{-1}$ ($\sigma =5\,$mJy\,beam$^{-1}$). The cross indicates the position of the phase tracking center of the interferometer. The $1''.8\x 1''.4$ clean beam is shown in the lower right box.} \label{fig:12co21_channel} \end{figure*} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[angle=270]{8337.f4}} \caption{Position velocity diagrams in $^{12}$CO(2--1) (left, contour levels 0.03\,Jy\,beam$^{-1}$) and $^{12}$CO(1--0) (right, contour levels 0.015\,Jy\,beam$^{-1}$). The horizontal coordinate is a distance offset (in arcsec) from the dynamical center along the line of nodes (p.a. 5\deg). The vertical coordinate is a radial velocity offset relative to 2470\,km\,s$^{-1}$. } \label{fig:posvel} \end{figure*} \subsection{Line ratios} For a quantitative analysis of the previous maps, we made maps of the ratios of the various integrated intensities, corrected for their respective primary beams. This correction makes the noise non-uniform through the ratio maps, especially in the $^{12}$CO(2--1) transition. Figures~\ref{fig:12_13} and \ref{fig:21_10} show the ratios $^{12}$CO(1--0)/$^{13}$CO(1--0) and $^{12}$CO(2--1)/$^{12}$CO(1--0) respectively. Each ratio map was made by smoothing the map with the higher resolution to that of the lower-resolution map. {\bf Ratio map $R_{12/13}=^{12}$\rm CO(1--0)/$^{13}$CO(1--0) :} The resolution is $\simeq 3''.1$. $^{13}$CO(1--0) is detected in the same places as $^{12}$CO(1--0), so the ratio can be studied in the entire CO nuclear disk. The average value is $<R_{12/13}>\simeq 9.5\pm 3.5$. The ratio is about 6 to 8 in the central zone (inside the two CO arcs), with the lowest value ($R\simeq 6$) near the center of {NGC$\,$1530}. The value is 11 to 15 in the arcs, with a maximum value of 15. The spatial distribution of dense gas ($n\simeq 10^4$\,cm$^{-3}$) is best shown in the HCN map (see Fig.~\ref{fig:molecules}). The ratio CO/HCN is 7 to 10 in the central ring of {NGC$\,$1530}\ (between the two arcs), while in the arcs this ratio is larger, in the range 14 to 30. The $^{12}$\rm CO(1--0)/$^{13}$CO(1--0) ratio thus seems to have the same characteristics as the CO/HCN ratio. \begin{figure*}[htb] \resizebox{\hsize}{!}{\includegraphics[angle=270]{8337.f5}} \caption{Ratio of integrated intensity $^{12}$CO(1--0)/$^{13}$CO(1--0)) in the central $25''\x 20''$ of {NGC$\,$1530}. The ratio was calculated with a $2\sigma$ threshold for each transition. The greyscale runs from 4 to 20 (white to black). The contour levels are from 5 to 20 by 3. Labels indicate levels 8 and 11. The beam is indicated by an ellipse in the lower left corner. The {\bf X} sign indicates the position of the dynamical center.} \label{fig:12_13} \end{figure*} {\bf Ratio map $R_{21/10}=^{12}$\rm CO(2--1)/$^{12}$CO(1--0) :} The resolution is $\simeq 1''.8$. The ratio can be studied in the entire disk with a high signal-to-noise ratio. The average ratio is $<R_{21/10}> \simeq 0.7\pm 0.2$. The ratio is $\simeq 1.0$ over a large region $4''\x 2''$ wide, with the dynamical center of the galaxy on the eastern edge of this region (see Fig.~\ref{fig:21_10}). The maximum value is $1.2$, at a position $3''$ west of the dynamical center. Between the two arcs, the ratio is generally $>0.7$. In the northern arc, the ratio is 0.4 to 0.7 while in the southern arc it is 0.5 to 1.1. \begin{figure*}[htb] \resizebox{\hsize}{!}{\includegraphics[angle=270]{8337.f6}} \caption{Ratio of integrated intensity $^{12}$CO(2--1)/$^{12}$CO(1--0) in the central $25''\x 20''$ of {NGC$\,$1530}. The ratio was calculated with a $3\sigma$ threshold for each transition. The greyscale runs from 0.3 to 1.2 (white to black). The contour levels are from 0.4 to 1.3 by $0.3$. Labels indicate levels 0.7 and $1.0$ (in white contour). The beam is indicated by an ellipse in the lower left corner. The {\bf X} sign indicates the position of the dynamical center.} \label{fig:21_10} \end{figure*} \subsection{Radio continuum maps} \begin{figure*} \resizebox{\hsize}{!}{\includegraphics[angle=270]{8337.f7}} \caption{$a)$ Left: $^{12}$CO(2--1) superposed on a grey scale map of the radio continuum at $\lambda= 20$\,cm. The grey scale runs from $0.18\simeq 2\sigma$ to 1.0\,mJy\,beam$^{-1}$). $b)$ Right: $^{12}$CO(2--1) superposed on a grey scale map of the radio continuum at $\lambda= 6$\,cm. The grey scale runs from $0.10\simeq 2\sigma$ to 0.42\,mJy\,beam$^{-1}$. The contours show the 6, 14 (black) and 18\,mJy\,beam$^{-1}$ (white) levels of the integrated intensity of $^{12}$CO(2--1). The cm-radio continuum lobes are indicated by ellipses in the lower left corners. The dynamical center is indicated by a {\bf X} sign. The cm-radio continuum maps are partially presented in Saikia et al. (1994), and were made kindly available by Dr.~A. Pedlar.} \label{fig:cm} \end{figure*} {NGC$\,$1530}\ was observed with the Very Large Array\footnote{The VLA is a facility of the National Radio Astronomy Observatory.} (VLA) in a snapshot mode. Saikia et al. (1994) show the 20\,cm emission map made with uniform weighting. Figure~\ref{fig:cm} shows the same data, at 20 and 6\,cm, in maps obtained with natural weighting, which allows maximum sensitivity. Superposed on the cm maps are a few contours of the $^{12}$CO(2--1) distribution. The beam sizes are $1''.6\x 1''.5$ (p.a. $-35\deg$) at 20\,cm and $1''.6\x 1''.3$ (p.a. $-32\deg$) at 6\,cm, both similar to the beam at CO(2--1). The emission peaks are $11\sigma$ at 20\,cm and $8\sigma$ at 6\,cm, so the detected features are not prominent in the maps. The 20\,cm distribution has two types of emission : a weak ($3\sigma$) emission over most of the molecular disk, and a number of compact components, unresolved by the interferometer. These components are mainly distributed along a ring, around a central cavity. The radius of this ring is about $3''$, i.e. 500\,pc. This ring coincides with the one obtained in the $^{12}$CO(2--1) and $^{12}$CO(1--0) lines. A strong component is detected $5''$ south of the dynamical center. It coincides well with a CO compact component. The 6\,cm map shows with a lower signal-to-noise ratio the same distribution as the 20\,cm map, i.e. a strong ring distribution. For the central $30''$ of {NGC$\,$1530}, integrated fluxes are 30\,mJy at 20\,cm and 8.6\,mJy at 6\,cm (integration threshold : $3\sigma$). The galaxy emits a total of 80\,mJy at 20\,cm and 37\,mJy at 6\,cm (Wunderlich et al. 1987, Condon et al. 1996). Thus $37\%$ and $23\%$ are emitted in the central $30''$ (i.e. 5\,kpc along the major axis) of {NGC$\,$1530}. The 500\,pc ring shares about $50\%$ of the central centimeter continuum emission, which is more than the $^{12}$CO(1--0) share of this ring, about 1/3 (RD 97). From the fluxes at 20\,cm and 6\,cm, we computed a spectral index of $-0.93$, which indicates that the synchrotron emission is predominant in these maps. There is a high star formation rate in the central part of {NGC$\,$1530}, giving rise to radio continuum emission via synchrotron emission from supernova remnants. These supernova remnants give rise to most of the compact sources in the cm maps. \section{Discussion} \subsection{Physical conditions of the gas} The previous maps reveal two main features for the molecular gas (also {\rm RD}\ 97) : {\em a)} Two intense arcs in both $^{12}$CO lines and with little HCN; {\em b)} A nuclear ring or spiral with strong HCN and $^{13}$CO(1--0). The radio continuum maps show a distribution similar to HCN, an intense ring and weak arcs. To discuss these results more quantitatively, we chose five regions in the nuclear disk of {NGC$\,$1530}, and calculated the various line ratios for these regions. Figure~\ref{fig:molecules} shows these regions on the $^{12}$CO(1--0) map. Table~\ref{tab:ratios} gives their coordinates and the line ratios, along with the intensity of $^{12}$CO(1--0) and fluxes from the cm maps convolved to a $2''.5$ gaussian for a better signal to noise ratio. We now relate the kinematics and the physical conditions of the gas in these regions. \medskip \begin{table*} \caption{Line ratios and cm fluxes in five regions of the center of {NGC$\,$1530}.} \begin{tabular}{c|c c c c c c c c c} \hline Region & R.A. $^a$ & Dec. & $I{\rm (CO)}$ $^b$ & $R_{21/10}$ $^c$ & $R_{12/13}$ & $R_{\rm CO/HCN}$ & $S_{20\,{\rm cm}}$ $^d$ & $S_{6{\rm cm}}$ & Spectral Index \\ & (s) & ($''$) & (K\,$\kms$) & & & & (mJy\,beam$^{-1}$) & (mJy\,beam$^{-1}$) &\\ \hline 1 & 26.0 & 45.5 & 183 & 0.91 & 6.9 & 8.0 & 1.87 & 0.78 & $-0.72$ \\ 2 & 27.0 & 42.0 & 206 & 0.77 & 6.9 & 13.0 & 1.77 & 0.63 & $-0.85$ \\ 3 & 28.0 & 49.0 & 273 & 0.67 & 9.9 & 28.8 & 0.78 & 0.30 & $-0.79$ \\ 4 & 25.5 & 38.0 & 104 & 0.82 & 8.9 & 14.1 & 1.09 & 0.51 & $-0.63$ \\ 5 & 24.7 & 40.5 & 168 & 0.76 & 14.9 & $>25.6$ & 0.52 & 0.12 & $-1.22$ \\ \hline \multicolumn{10}{l}{{\scriptsize \it a} Coordinates (J2000): $04^{\rm h} 23^{\rm m}$; $+75\deg 17'$}\\ \multicolumn{10}{l}{{\scriptsize \it b} $^{12}$CO(1--0) integrated intensity}\\ \multicolumn{10}{l}{{\scriptsize \it c} Line ratios $^{12}$CO(2--1)/$^{12}$CO(1--0), $^{12}$CO(1--0)/$^{13}$CO(1--0) and $^{12}$CO(1--0)/HCN(1--0)}\\ \multicolumn{10}{l}{{\scriptsize \it d} 20\,cm, 6\,cm fluxes in $2''.5$ diameter lobes}\\ \end{tabular} \label{tab:ratios} \end{table*} Regions 2, 3 and 5 correspond to local maxima of $^{12}$CO(1--0) and $^{12}$CO(2--1) emission. Region 1 is a local maximum of $^{12}$CO(2--1) and HCN(1--0) while region 4 corresponds to a local minimum of $^{12}$CO(1--0) and $^{12}$CO(2--1). These regions are typical of the differing conditions in the nuclear disk. We have compared the location of these five regions with a CO(1--0) velocity map from {\rm RD}\ 97. Regions 1 and 2 correspond to a nearly circular rotation of the gas, as they are at the contact point between the arcs and the nuclear ring. They are close to the dynamical center, at a radius of 0.7\,kpc. Regions 3 and 4, at radii of 1.7\,kpc, correspond to transition points in the kinematics, a transition between infall motions (in the CO arcs) and circular rotation (in the ring). Region 5 is further out in the southwest arc, at a 3\,kpc radius. The motion of this region has an infall component of $70\kms$. This infall motion is associated with the density wave of the arc ({\rm RD}\ 97). The molecular ratios seem normal in the nuclear disk of {NGC$\,$1530}. We compare the central kpc of {NGC$\,$1530}\ to the center of the spiral galaxy IC342 which probably contains a weak bar (see Downes et al. 1992, Wright et al. 1993). IC\,342 is one of the rare galaxies which has been observed in several different molecules with interferometers. IC\,342 shows straight regular $^{12}$CO(1--0) lanes (Wright et al. 1993, their Fig.~2b) that are curved near the nucleus, similarly to {NGC$\,$1530}. The lanes emit in $^{13}$CO(1--0) (Wright et al. 1993, their Fig.~2b) but not in HCN(1--0) (Downes et al. 1992, their Fig.~1); in these two transitions one finds five 50\,pc diameter clumps at the places where the lanes curve toward the nucleus. These clumps are not prominent in $^{12}$CO(1--0). $R_{12/13}$ is 4.4 in these clumps and $\geq 20$ in the CO lanes. Similarly, $R_{\rm CO/HCN}$ is 7 in the clouds and $\geq 20$ in the CO lanes. Therefore the results seem similar for both galaxies, with ratios $R_{12/13}$ and $R_{\rm CO/HCN}$ lower near the nucleus ($r<100$\,pc for IC342, $r<700$\,pc for NGC\,1530). The places where the CO lanes become a nuclear ring or spiral show large emissivity in $^{13}$CO(1--0) and HCN(1--0), which can be interpreted as the sign of a high concentration of dense gas (see, e.g., Downes et al. 1992, Mauersberger \& Henkel 1993). \begin{table*} \caption{Results from an escape probability analysis for the regions of Table~\ref{tab:ratios}.} \begin{tabular}{c| c c c c c c} \hline Region & $n(\rm H_2)$ $^a$ & $T_{\rm kin}$ $^b$ & $T_{\rm b}^{\rm obs}$ $^c$ & $T_{\rm b}^{\rm th}$ $^d$ & $f$ $^e$ & $M(\rm H_2)/\mathit L_{\rm CO}$ $^f$ \\ & ($10^2\x$cm$^{-3}$) & (K) & (K) & (K) & & \\ \hline 1 & 8.2 & 51 & 2.7 & 32.3 & 0.08 & 1.2\\ 2 & 4.7 & 20 & 2.6 & 12.8 & 0.2 & 2.3\\ 3 & 3.1 & 18 & 3.2 & 9.8 & 0.3 & 2.0\\ 4 & 4.5 & 40 & 1.5 & 20.6 & 0.07 & 1.1\\ 5 & 2.1 & 85 & 2.1 & 16 & 0.13 & 0.35\\ \hline \multicolumn{7}{l}{{\scriptsize \it a} H$_2$ density}\\ \multicolumn{7}{l}{{\scriptsize \it b} Model kinetic temperature}\\ \multicolumn{7}{l}{{\scriptsize \it c} Observed $^{12}$CO(1--0) brightness temperature}\\ \multicolumn{7}{l}{{\scriptsize \it d} Model $^{12}$CO(1--0) brightness temperature}\\ \multicolumn{7}{l}{{\scriptsize \it e} Area filling factor of molecular clouds}\\ \multicolumn{7}{l}{{\scriptsize \it f} Conversion factor (in \mo\,(K\,$\kms$\,pc$^2$)$^{-1}$) computed}\\ \multicolumn{7}{l}{from $M(\rm H_2)/\mathit L_{\rm CO} = 2.1\,n(\rm H_2)^{1/2}/T_{\rm b}$(CO(1--0))}\\ \end{tabular} \label{tab:lvg} \end{table*} We supposed that the $^{12}$CO and $^{13}$CO transitions are emitted by the same masses of gas. Therefore we could run escape probability models, to reproduce the observed line ratios of CO. We assumed the following values for the abundance ratios : [$^{12}$CO]/[H$_2$]$=10^{-4}$ and [$^{12}$CO]/[$^{13}$CO]$=60$. We assumed a velocity gradient of $1\,$km\,s$^{-1}$\,pc$^{-1}$. Table~\ref{tab:lvg} displays the results of the best fitting model. We deduced from this model the expected H$_2$ density and kinetic temperature of the emitting gas for each region. Then we could derive the theoretical brightness temperature and obtain a filling factor by computing the ratio $T_{\rm b}^{\rm obs}/T_{\rm b}^{\rm th}$. We computed also for each one of these 5 regions the CO to H$_2$ conversion factor, using the formula $M(\rm H_2)/\mathit L_{\rm CO}= 2.1\,n(\rm H_2)^{1/2}/T_{\rm b}$(CO(1--0)), in units of \mo\,(K\,$\kms$\,pc$^2$)$^{-1}$ (Radford et al. 1991). This formula is valid for an ensemble of virialized molecular clouds, which is the case here. The derived kinetic temperatures are standard values for the molecular gas, between 20 and 90\,K. This kinetic temperature is weakly constrained by the escape probability calculations, and the error bars of the measured ratios do not allow a precise derivation. The H$_2$ density is more tightly constrained. The gas density is greater near the center ($\gsim 5\E2 $\,cm$^{-3}$) than in the arms (2 to $3\E2 $). Region 4 has a greater density than the equivalent region 3. Regions with greater density (1, 2, 4) show a stronger HCN emission than the ones with a lower density, confirming the association of HCN emission with dense gas (Mauersberger \& Henkel 1993). The derived filling factors are loosely constrained, as are the theoretical CO(1--0) brightness temperatures, depending on the kinetic temperatures. The average value is $<f>\simeq 0.15$. These low values reveal the clumpy nature of the molecular gas, contained in many small molecular clouds, unresolved with the $\simeq 500\,$pc beam of the interferometer. The derived values for the conversion factor have a average of $<M(\rm H_2)/\mathit L_{\rm CO}>\simeq 1.4$\,\mo\,(K\,$\kms$\,pc$^2$)$^{-1}$. This is lower than the average value for the giant molecular clouds of our own galaxy ($M(\rm H_2)/\mathit L_{\rm CO} \simeq 4.8$\,\mo\,(K\,$\kms$\,pc$^2$)$^{-1}$, Sage \& Isbell 1991). Region 5 has a conversion factor significantly lower than the others, possibly due to different excitation conditions for the molecular transitions. We found in the central kiloparsec of {NGC$\,$1530}\ a\\ $M(\rm H_2)/\mathit L_{\rm CO}$ ratio about three times lower than the standard galactic value of Sage \& Isbell (1991). It is unlikely that this conversion factor is universal. For the inner 14\,kpc diameter region of NGC\,891, Gu\'elin et al. (1993) found a ratio 3 times lower than the standard galactic value. For the innermost 11\,kpc of M51, Gu\'elin et al. (1995) found a ratio 4 times lower than the standard galactic value. Our results are similar to the results found by Gu\'elin et al. (1993, 1995). In the inner 1200 pc of the Milky Way, Dahmen et al. (1998) found a factor of 10 discrepancy relative to the standard Galactic value. We do not find comparable results in the inner kiloparsec of {NGC$\,$1530}. \subsection{Star formation and dense gas} Helou et al. (1985) found a proportionality between the far infrared flux and the 20\,cm flux for disks of spiral galaxies. This proportionality indicates that the {\em global} star formation rate and the {\em global} 20\,cm flux are linearly correlated. We assumed a linear correlation between the {\em local} star formation rate and the {\em local} 20\,cm flux density. Thus for each of the 5 regions we computed its local star formation rate $SFR^{\rm loc}$ in a $2''.5$ beam by the following formula, with $S^{\rm tot}_{20\rm cm} =80\,$mJy (total 20\,cm flux from {NGC$\,$1530}, from Wunderlich et al. 1987, Condon et al. 1996) and $SFR^{\rm tot} = 2.4\mo$\,yr$^{-1}$ \[ SFR^{\rm loc} = \frac{S^{\rm loc}_{20\rm cm}}{S^{\rm tot}_{20\rm cm}} SFR^{\rm tot} \] To compare with the dense gas distribution, we calculated the amount of H$_2$ gas in dense phase present in each one of these five regions. We used for that the relation existing between the mass of dense gas and the velocity integrated HCN intensity. This relation is derived from HCN radiative transfer solutions (Solomon et al. 1992). It can be written as $M_{\rm HCN}(\rm H_2)/{\it L}_{\rm HCN}\simeq 20\,$\mo\,(K\,$\kms$\,pc$^2$)$^{-1}$, where $M_{\rm HCN}(\rm H_2)$ is the mass of H$_2$ at a density $\simeq 10^4\,$cm$^{-3}$ as traced by HCN. From the HCN fluxes of Fig.~\ref{fig:molecules} and this relation, we computed the molecular gas masses of Table~\ref{tab:sfr}. The comparison of $SFR^{\rm loc}$ and $M_{\rm HCN}(H_2)$ show that the star formation rate seems correlated with the amount of available dense gas. The region 4 shows a higher $SFR^{\rm loc}$ and a higher amount of dense gas than regions 3 and 5, even though region 4 is included in an arc, like the other two regions. Region 5 shows a very low level of star formation, consistent with its cm wavelength spectral index, $-1.2$ (see Table~\ref{tab:ratios}), which indicates the absence of thermal component in the cm emission. \begin{table} \caption{Star formation rates and masses of dense gas (n({H$_2$)}$\simeq 10^4$\,cm$^{-3}$) in the five regions of Table~\ref{tab:ratios}.} \begin{tabular}{c| c c} \hline Region & $SFR^{\rm loc}$ $^a$ & $M_{\rm HCN}$(H$_2$) $^b$ \\ & ($10^{-2}\mo$\,an$^{-1}$) & ($10^7\mo$) \\ \hline 1 & 6 & 16\\ 2 & 5 & 11\\ 3 & 2 & 4\\ 4 & 3 & 8\\ 5 & 1 & 5\\ \hline \multicolumn{3}{l}{{\scriptsize \it a} Local star formation rates from 20\,cm flux in a $2''.5$ beam}\\ \multicolumn{3}{l}{{\scriptsize \it b} Dense gas masses from HCN integrated fluxes in a $3''.5$}\\ \multicolumn{3}{l}{beam from $M_{\rm HCN}(\rm H_2)/{\it L}_{\rm HCN}\simeq 20\,$\mo\,(K\,$\kms$\,pc$^2$)$^{-1}$}\\ \end{tabular} \label{tab:sfr} \end{table} \section{Conclusion} We studied the inner molecular disk of the barred spiral galaxy {NGC$\,$1530}\ by means of millimeter interferometry. The main conclusions of this paper are :\\ --- The $^{12}$CO(2--1) intensity map obtained with a $1''.6$ resolution is very similar to the $^{12}$CO(1--0) distribution, with a nuclear ring, and around it, two bright curved arcs. \\ --- The $^{13}$CO(1--0) intensity map synthesized with a $3''.1$ beam is rather similar to the $^{12}$CO(1--0) map, with weaker curved arcs.\\ --- The average ratios are $R_{12/13}=9$ and $R_{21/10}=0.7$, comparable to other spiral galaxies. $R_{12/13}$ shows a general decrease from the central ring ($R_{12/13}\simeq 6-8$) to the curved arcs($R_{12/13}\simeq 11-15$). $R_{21/10}$ shows no such systematic trend. The ratio is highest ($\simeq 1.2$) in a region close to the dynamical center.\\ --- The VLA cm maps show a ring very similar to the $^{12}$CO ring, but little or no emission in the curved arcs.\\ --- An escape probability analysis reproduces well the line ratios in five regions of the disk. We could derive the kinetic temperature and the density of the gas, together with filling factor. The gas density seems higher in the nuclear ring than in the curved arcs, except for an isolated region in the south-western arc.\\ --- The conversion factor deduced from this model is\\ $ M(\rm H_2)/\mathit L_{\rm CO} \simeq 1.4$\,\mo\,(K\,$\kms$\,pc$^2$)$^{-1}$, three times lower than the standard galactic value of $4.8$\,\mo\,(K\,$\kms$\,pc$^2$)$^{-1}$.\\ --- The star formation rate as calculated from the 20\,cm flux and the mass of dense gas deduced from the HCN flux are spatially correlated.\\ \begin{acknowledgements} We thank the technical staff of the IRAM interferometer for the remarkable job they do. We thank Dr. Alan Pedlar for making available the VLA data. We thank the referee Dr. Vogel for helpful comments. \end{acknowledgements}
\section{Introduction.} \-In this note we consider a sine-Gordon soliton in the presence of a periodic potential. The soliton is fast moving, thus relativistic effects shall be taken into account. In particular we expect the total energy of the system to be divided into two contributions, one equal to $M/\sqrt{1-B^{2}}$ corresponding to a soliton of mass $M$ moving at velocity $B$ and another pertaining to the field of fluctuations about the soliton solution. A similar separation should hold for the total momentum of the system where the term $MB/\sqrt{1-B^{2}}$ is expected to appear. Clearly expanding around the usual static solution (\ref{1}) is certainly not good enough:\ A fast moving soliton suffers Lorentz contraction and the Lorentz factor should appear in the classical solution. \- The soliton interacts both with the fluctuating field and with the external potential; it will prove both natural and fruitful to treat the velocity $B$ as a dynamical variable. This is essential if recoil effects are to be taken into account. A formalism along these lines is presented in \cite{1}; however the present note is meant to be self-contained. Note that (a) the model considered in \cite{1} was a pure $\phi ^{4}$ theory with no external potential (the work focused on the interaction of the soliton with the fluctuations), (b) the emphasis was on the quantization of the theory. A model of a soliton in an external potential is presented in \cite{2} but there the aim was to examine the motion of the soliton, radiation effects being disregarded. The problem examined in the present note has been looked at in \cite{4}; here we put an emphasis on a relativistic treatment of the problem. \section{\noindent Basics.} The simple sine-Gordon Lagrangean is given by (for a review see \cite{3}, \cite{8}) \begin{equation} L=\int dx^{\prime }\left\{ \frac{1}{2}\left( \frac{\partial \phi }{\partial t^{\prime }}\right) ^{2}-\frac{1}{2}\left( \frac{\partial \phi }{\partial x^{\prime }}\right) ^{2}-U_{sG}(\phi )\right\} \label{2} \end{equation} \noindent where \begin{equation} U_{sG}(\phi )=\frac{\alpha }{\beta ^{2}}\left( 1-\cos (\beta \phi )\right) \label{6} \end{equation} \noindent (Lab frame coordinates are denoted by $(t^{\prime },x^{\prime }))$% . We make the choice \begin{equation} \alpha =\beta =1 \label{8} \end{equation} \noindent and then \begin{equation} U_{sG}(\phi )=1-\cos \phi \label{13} \end{equation} \noindent Eventually we will consider the soliton in an external perturbing potential of the form \cite{4} \begin{equation} U(\phi )=U_{sG}(\phi )\left( 1+V(x^{\prime })\right) \label{25} \end{equation} \begin{equation} V(x^{\prime })=\varepsilon \cos (k_{0}x^{\prime }) \label{v} \end{equation} where $\varepsilon $ is a small quantity. The equation of motion resulting from (\ref{2}) is \begin{equation} \frac{\partial ^{2}\phi }{\partial t^{\prime 2}}-\frac{\partial ^{2}\phi }{% \partial x^{\prime 2}}+U_{sG}^{\prime }(\phi )=0 \label{4} \end{equation} \noindent The static version \begin{equation} -\frac{d^{2}\phi }{dx^{\prime 2}}+U_{sG}^{\prime }(\phi )=0 \label{4b} \end{equation} \noindent admits (amongst others) the solution \begin{equation} \Phi _{c}(x^{\prime })=4\arctan (e^{x^{\prime }}) \label{1} \end{equation} \noindent where $\Phi _{c}\rightarrow 0$ for $x^{\prime }\rightarrow -\infty $ and $\Phi _{c}\rightarrow 2\pi $ for $x^{\prime }\rightarrow \infty $. The mass of the soliton is quite generally given by \begin{equation} M=\int dx^{\prime }\left\{ \frac{1}{2}\left( \frac{\partial \Phi _{c}}{% \partial x^{\prime }}\right) ^{2}+U_{sG}(\Phi _{c})\right\} =\int dx^{\prime }\left( \frac{\partial \Phi _{c}}{\partial x^{\prime }}\right) ^{2} \label{19} \end{equation} \noindent and for the particular choice (\ref{8}) $M=8.$ Fluctuations around the classical solution satisfy the equation (expand (\ref{4}) around $\Phi _{c}$) \begin{equation} -\frac{d^{2}f}{dx^{\prime 2}}+U_{sG}^{\prime \prime }(\Phi _{c})f=\omega ^{2}f \label{5} \end{equation} \noindent with $\omega $ the corresponding frequency. Note that the space derivative of the kink \begin{equation} \Phi _{c}^{\prime }(x^{\prime })=\frac{2}{\cosh x^{\prime }} \label{zero} \end{equation} \noindent is a solution corresponding to zero frequency (the zero mode) (in the last equation we used (\ref{1})). This simply reflects the fact that a translated kink $\Phi _{c}(x^{\prime }+a)$ is also a static solution. The normal modes corresponding to a wavevector $k$ (not to be confused with $% k_{0}$ of (\ref{v}) that is a characteristic of the potential) are given in \cite{4} : \begin{equation} \frac{1}{\omega (k)\sqrt{2\pi }}e^{ikx^{\prime }}\left( k+i\tanh x^{\prime }\right) \label{14} \end{equation} \noindent where $\omega =\sqrt{k^{2}+\mu ^{2}}$ , and $\mu =1$ with the choice (\ref{8}). We shall write the normal modes in the form \begin{equation} f(k,x)=\tilde{A}(k,x^{\prime })e^{ikx^{\prime }} \label{14b} \end{equation} \noindent where the prefactor $\tilde{A}(k,x^{\prime })$ can be read off (% \ref{14}) modulo a norming factor; we will work with energy normalized wavefunctions \[ \int dxf^{*}(k_{1},x)f(k_{2},x)=\delta (\omega _{1}-\omega _{2}) \] The Hamiltonian is \begin{equation} H=\int dx\left\{ \frac{1}{2}\left( \frac{\partial \phi _{c}}{\partial x^{\prime }}\right) ^{2}+\frac{\pi _{\phi }^{2}}{2}+U_{sG}(\phi )\right\} \label{15} \end{equation} \noindent The fields are expanded in terms of the zero mode and the travelling modes \begin{equation} \phi (x^{\prime },t^{\prime })=\Phi _{c}(x^{\prime })+q_{z}\Phi _{c}^{\prime }(x^{\prime })+\frac{1}{\sqrt{4\pi }}\int \frac{d\omega }{\sqrt{\omega }}% \left\{ \tilde{a}^{\dagger }(\omega )f(k,x^{\prime })e^{i\omega t^{\prime }}+% \tilde{a}(\omega )f^{*}(k,x^{\prime })e^{-i\omega t^{\prime }}\right\} \label{16} \end{equation} \begin{equation} \pi _{\phi }(x^{\prime },t^{\prime })=\frac{p_{z}}{M}\Phi _{c}^{\prime }(x^{\prime })+\frac{1}{\sqrt{4\pi }}\int d\omega \sqrt{\omega }\left\{ \tilde{a}^{\dagger }(\omega )f(k,x^{\prime })e^{i\omega t^{\prime }}-\tilde{a% }(\omega )f^{*}(k,x^{\prime })e^{-i\omega t^{\prime }}\right\} \label{17} \end{equation} \noindent with the commutation relations \begin{equation} \left[ q_{z},p_{z}\right] =1 \label{18} \end{equation} \[ \left[ \tilde{a}(\omega ),\tilde{a}^{\dagger }(\omega )\right] =\delta (\omega -\omega ^{\prime }) \] \noindent $\phi ,\Pi _{\phi }$ obey the standard canonical commutation relations. (NB\ In what follows and for the sake of brevity integration over $\omega $ will also stand for summation over the two directions of incidence.) Notice that $p_{z}$ appears in the Hamiltonian in the form $% {\displaystyle {p_{z}^{2} \over 2M}} $ through the $% {\displaystyle {\pi _{\phi }^{2} \over 2}} $ term. Thus $p_{z}$ can be identified as the momentum of the soliton. The fact that the kinetic energy of the soliton appears in the nonrelativistic form is connected to the fact that we consider fluctuations about the zero-velocity solution (\ref{1}). Clearly the inclusion of relativistic effects requires expansion about a boosted version of (\ref{1}). Neglecting the interaction with fluctuations the motion of the kink at a nonrelativistic speed $u$ corresponds to \begin{equation} q_{z}=ut^{\prime } \label{30} \end{equation} \begin{equation} p_{z}=Mu \label{31} \end{equation} \noindent compatible with $\pi _{\phi }=\partial \phi /\partial t^{\prime }.$ We introduce the fluctuating field \begin{equation} \chi (x^{\prime },t^{\prime })\equiv \phi (x^{\prime },t^{\prime })-\Phi _{c}(x^{\prime }) \label{20} \end{equation} \noindent and split the Hamiltonian (\ref{15}) into a part quadratic in $% \chi $ and another part involving higher order terms. Then the quadratic part (diagonalized by the normal modes) does not involve $q_{z}$. Thus the momentum of the kink (identified with $p_{z}$) is conserved; it can change only through interactions either with the fluctuating field or with an external potential. \section{\noindent \noindent The moving soliton.} \subsection{The Hamiltonian formalism} In order to treat a fast moving soliton we employ the strategy used in \cite {1} and in a somewhat different context in \cite{5}, namely we go over to a non-inertial frame comoving with the soliton. The definition of the position of the soliton is of course not a straightforward affair and it will occupy us later on. The new coordinates are defined by \begin{equation} x^{\prime }=x+X(t)\text{ , }t^{\prime }=t\text{ , }B\equiv \dot{X}(t) \label{21} \end{equation} \noindent We redefine the field so that \begin{equation} \Phi (x)=\phi (x+X) \label{24} \end{equation} \noindent (Transformations (\ref{21}), (\ref{24}) coincide with the translation transformation implemented quantum-mechanically in \cite{6}, section 3.1). Thus \begin{equation} \frac{\partial \phi }{\partial x^{\prime }}=\frac{\partial \Phi }{\partial x}% \text{ , }\frac{\partial \phi }{\partial t^{\prime }}=\frac{\partial \Phi }{% \partial t}-B\frac{\partial \Phi }{\partial x} \label{22} \end{equation} \noindent Lagrangean (\ref{2}) with the potential (\ref{25}) can be readily written in terms of the new coordinates \begin{equation} L=\int dx\left\{ \frac{1}{2}\left( \frac{\partial \Phi }{\partial t}-B\frac{% \partial \Phi }{\partial x}\right) ^{2}-\frac{1}{2}\left( \frac{\partial \Phi }{\partial x}\right) ^{2}-U_{sG}(\Phi (x))(1+V(x+X))\right\} \label{230} \end{equation} \noindent The momentum $\Pi $ conjugate to $\Phi $ is \begin{equation} \Pi =\frac{\partial \Phi }{\partial t}-B\frac{\partial \Phi }{\partial x} \label{26} \end{equation} \noindent whereas the momentum $P$ conjugate to $X$ satisfies the constraint \begin{equation} P+\int dx\Pi \frac{\partial \Phi }{\partial x}=0 \label{27} \end{equation} \noindent The constraint reflects the arbitrariness in what we mean by $X$ (i.e. the freedom to perform coordinate transformations). The canonical Hamiltonian $H_{c}=p\dot{q}-L$ is \begin{equation} H_{c}=\int dx\left\{ \frac{\Pi ^{2}}{2}+\frac{1}{2}\left( \frac{\partial \Phi }{\partial x}\right) ^{2}+U_{sG}(\Phi (x))(1+V(x+X))\right\} \label{28} \end{equation} \noindent The numerical value of $H_{c}$ coincides with the total energy of the system. The total Hamiltonian (i.e. the quantity that yields the equations of motion) consists of the canonical Hamiltonian $H_{c}$ plus the constraint multiplied by an arbitrary Lagrange multiplier \cite{7} \begin{equation} H_{c}+B\left( P+\int dx\Pi \frac{\partial \Phi }{\partial x}\right) \label{29} \end{equation} \noindent The identification of the Lagrange multiplier $B$ as the velocity follows immediately by commuting $X$ with (\ref{29}) and taking into account the fact that $X$ and $P$ are canonically conjugate. Since the velocity changes as a result of interactions with the fluctuations and/or the external potential it is natural to elevate it to the status of a dynamical variable and introduce its conjugate momentum $p_{B}$% \begin{equation} \left[ B,p_{B}\right] =1 \label{33} \end{equation} \noindent We take $B,p_{B}$ to commute with all other variables $\Phi ,\Pi ,X,P$. We require $p_{B}$ to vanish as a constraint and introduce the (final) total Hamiltonian \begin{equation} H=H_{c}+B\left( P+\int dx\Pi \frac{\partial \Phi }{\partial x}\right) +ap_{B} \label{32} \end{equation} \noindent where we added to the Hamiltonian (\ref{29}) constraint $p_{B}$ multiplied by a Lagrange multiplier $a$. Commuting $B$ with $H$ we get \begin{equation} \dot{B}=a \label{34} \end{equation} \noindent Thus $a$ is identified as the acceleration of the kink. Requiring $% p_{B}$ to be conserved we end up with the original constraint (\ref{27}). The situation is reminiscent of what happens in electrodynamics. The momentum conjugate to the scalar potential (the latter being the analog of $% B $ in the present case) vanishes and the requirement that the constraint be conserved leads to Gauss's law (the analog of (\ref{27})) as a secondary constraint. The gauge invariance reflects the freedom of performing coordinate transformations of the type (\ref{21}). It will be lifted when we impose a subsidiary condition (thus implicitly defining the position of the soliton and hence the meaning of a comoving frame). If we neglect for the moment the effect of the perturbation $V$, the field equation resulting from (\ref{230}) (or (\ref{28}) and (\ref{32})) for constant velocity $B$ is \begin{equation} \frac{\partial ^{2}\Phi }{\partial t^{2}}-\frac{\partial ^{2}\Phi }{\partial x^{2}}+U_{sG}^{\prime }(\Phi )-2B\frac{\partial ^{2}\Phi }{\partial x\partial t}+B^{2}\frac{\partial ^{2}\Phi }{\partial x^{2}}=0 \label{351} \end{equation} \noindent A $t$ independent solution is trivially found to be \begin{equation} \Phi _{c}\left( \frac{x}{\sqrt{1-B^{2}}}\right) \label{352} \end{equation} \noindent \noindent where $\Phi _{c}$ is the same as (\ref{1}) (i.e. the space variable is simply rescaled by the Lorentz factor). It is natural to expand the fields $\Phi ,\Pi $ about the classical solution (\ref{352}) introducing the fluctuating fields $\chi $ and $\pi $ (the latter not to be confused with $\pi _{\phi }$ of section 2). (NB\ in what follows $\Phi _{c}^{^{\prime }}$ will refer to the derivative of $\Phi _{c}$ with respect to its argument and not just with respect to $x.)$ \begin{equation} \Phi (x,t)=\Phi _{c}\left( \frac{x}{\sqrt{1-B^{2}}}\right) +\chi (x,t) \label{35} \end{equation} \begin{equation} \Pi (x,t)=-\frac{B}{\sqrt{1-B^{2}}}\Phi _{c}^{^{\prime }}\left( \frac{x}{% \sqrt{1-B^{2}}}\right) +\pi (x,t) \label{36} \end{equation} \noindent The existence of the first term in (\ref{36}) is a consequence of the $\partial /\partial x$ term in (\ref{26}). It follows from (\ref{35}), (% \ref{36}) that $\chi ,\pi $ satisfy canonical commutation relations \begin{equation} \left[ \chi (x,t),\pi (x^{\prime },t)\right] =1 \label{39} \end{equation} \noindent It is important however to realize that $p_{B}$ does not commute with $\chi $ and $\pi $. Given that $p_{B}$ commutes with $\Phi ,\Pi $ it follows immediately from (\ref{35}), (\ref{36}) that \begin{equation} \left[ \chi (x,t),p_{B}\right] =-\Phi _{c}^{\prime }\left( \frac{x}{\sqrt{% 1-B^{2}}}\right) \frac{d}{dB}\left( \frac{1}{\sqrt{1-B^{2}}}\right) \label{37} \end{equation} \begin{equation} \left[ \pi (x,t),p_{B}\right] =\Phi _{c}^{^{\prime }}\left( \frac{x}{\sqrt{% 1-B^{2}}}\right) \frac{d}{dB}\left( \frac{B}{\sqrt{1-B^{2}}}\right) +\frac{% B^{2}}{\left( 1-B^{2}\right) ^{2}}\Phi _{c}^{^{\prime \prime }}\left( \frac{x% }{\sqrt{1-B^{2}}}\right) \label{38} \end{equation} We turn to the field momentum and express it in terms of the static solution and the fluctuations (and use (\ref{19})): \begin{equation} \int dx\Pi \frac{\partial \Phi }{\partial x}=-\frac{MB}{\sqrt{1-B^{2}}}+\int dx\pi \frac{\partial \chi }{\partial x}+\frac{1}{\sqrt{1-B^{2}}}\int dx\left\{ \Phi _{c}^{^{\prime }}\pi +\frac{B}{\sqrt{1-B^{2}}}\Phi _{c}^{^{\prime \prime }}\chi \right\} \label{41} \end{equation} \noindent It is gratifying that in the absence of fluctuations we get the expected relativistic expression for the kink's momentum. \subsection{The subsidiary condition.} We have to define what we mean by the position of the soliton or in other words give a definition of the comoving frame. This is done through the imposition of a constraint (a subsidiary condition) that lifts the freedom under coordinate transformations and assigns physical meaning to the variable $X.$ The choice of the suitable constraint is dictated by three criteria each one having its own physical motivation. It is not obvious at the outset that all three can be simultaneously satisfied. The fact that they can adds physical appeal to the separation of the total field $\Phi $ to a soliton part and a fluctuating part. The first criterion stipulates that in the presence of fluctuations the total momentum splits neatly to the soliton and fluctuation momenta given respectively by minus the first and second terms in (\ref{41}) and that the term in braces vanishes. In other words the constraint that lifts the gauge invariance is \begin{equation} C\equiv \int dx\left\{ \Phi _{c}^{^{\prime }}\pi +\frac{B}{\sqrt{1-B^{2}}}% \Phi _{c}^{^{\prime \prime }}\chi \right\} =0 \label{420} \end{equation} \noindent The fact that $C$ does not contribute numerically in (\ref{41}) does not mean that it should be discarded; it may well contribute to the equations of motion since the LHS\ of (\ref{41}) appears in the total Hamiltonian (\ref{32}). It turns out that it does not contribute; see the remark following (\ref{400e}). As a second criterion we require that the $energy$ $H_{c}$ of the system does not contain any terms linear in the fluctuations. In other words that the same state of affairs as in the case of the total momentum prevails. It turns out that this requirement is satisfied; see (\ref{400e}) and the remark following it. The third criterion stipulates that constraint $C$ commute with the quadratic part $H_{0}$ (\ref{50}) of the total Hamiltonian. This motivation is linked to the acceleration of the kink. The acceleration will be determined in the next subsection (eqn (\ref{451a})) by requiring that constraint $C$ commute with the total Hamiltonian (\ref{32}), and it has two physical origins:\ (i) The interaction of the kink with the fluctuations, i.e. terms in the Hamiltonian beyond the quadratic in expansion (\ref{400c}% ); (ii) The interaction with the external potential $V$. Thus in the absence of (i) and (ii) the acceleration ought to vanish, i.e. $C$ should commute with $H_{0}$ (\ref{50}). This is shown in (\ref{51b}) below. We substitute expansions (\ref{35}), (\ref{36}) in $H_{c}$ (\ref{28}) while expanding $U_{sG}(\Phi )$ up to terms quadratic in $\chi $. Use of (\ref{19}% ) and straight algebra yields \begin{eqnarray} H_{c} &=&\frac{M}{\sqrt{1-B^{2}}}+ \nonumber \\ &&+\int dx\left\{ \frac{1}{2}\left( \frac{\partial \chi }{\partial x}\right) ^{2}+\frac{\pi ^{2}}{2}+\frac{1}{2}U_{sG}^{\prime \prime }\left( \Phi _{c}\right) \chi ^{2}+V\text{-independent higher orders in }\chi \right\} + \nonumber \\ &&+\left\{ V(x+X)U_{sG}\left( \Phi _{c}\right) +V\text{-dependent higher orders in }\chi \right\} + \label{400c} \\ &&+\int dx\left\{ -\frac{B}{\sqrt{1-B^{2}}}\Phi _{c}^{^{\prime }}\pi -\frac{1% }{1-B^{2}}\Phi _{c}^{^{\prime \prime }}\chi +U_{sG}^{^{\prime }}\left( \Phi _{c}\right) \chi \right\} \nonumber \end{eqnarray} \noindent Thus $H_{c}$ naturally splits to a number of contributions: The first term coincides with the covariant expression for the energy of a particle. The next three terms are quadratic in the fluctuations and enter in the calculation of the normal modes of the system. There also are terms describing the interaction of the kink with the fluctuations and/or the external potential. The last term in braces linear in fluctuations can be written using (\ref{4b}) in the form \begin{equation} -\frac{B}{\sqrt{1-B^{2}}}\int dx\left\{ \Phi _{c}^{\prime }\pi +\frac{B}{% \sqrt{1-B^{2}}}\Phi _{c}^{\prime \prime }\chi \right\} \label{400e} \end{equation} \noindent Hence the second criterion spelt out in the beginning of this subsection is satisfied: the energy does not depend linearly on the fluctuations. Notice that (\ref{400e}) cancels exactly with the last term in (\ref{41}) when they are both substituted in (\ref{32}). Thus the total Hamiltonian $H$ is written in terms of $\chi ,\pi $ in the form \begin{eqnarray} H &=&M\sqrt{1-B^{2}}+ \nonumber \\ &&+\int dx\left\{ \frac{1}{2}\left( \frac{\partial \chi }{\partial x}\right) ^{2}+\frac{\pi ^{2}}{2}+\frac{1}{2}U_{sG}^{\prime \prime }\left( \Phi _{c}\right) \chi ^{2}+V\text{-independent higher orders in }\chi \right\} + \nonumber \\ &&+\left\{ V(x+X)U_{sG}\left( \Phi _{c}\right) +V\text{-dependent higher orders in }\chi \right\} + \label{hkhi} \\ &&+B\int dx\pi \frac{\partial \chi }{\partial x}+ap_{B}+BP \nonumber \end{eqnarray} \noindent The first terms in (\ref{41}) and (\ref{400c}) combine to yield the first term in (\ref{hkhi}) above. This latter term yields the frequency exponential $\exp (-M\sqrt{1-B^{2}}t)$ in the soliton wavefunction and is the relativistic generalization of the phase $\exp (+imB^{2}t/2)$ (with an apparently wrong sign) derived in \cite{9} for a point particle. We can now write down the part $H_{0}$ of the total Hamiltonian (\ref{hkhi}) quadratic in the fluctuations: \begin{equation} H_{0}=M\sqrt{1-B^{2}}+\int dx\left\{ \frac{1}{2}\left( \frac{\partial \chi }{% \partial x}\right) ^{2}+\frac{\pi ^{2}}{2}+\frac{1}{2}U_{sG}^{\prime \prime }\left( \Phi _{c}\right) \chi ^{2}\right\} +B\int dx\pi \frac{\partial \chi }{\partial x} \label{50} \end{equation} \noindent The equation of motion describing free fluctuations derivable from the above is \begin{equation} \frac{\partial ^{2}\chi }{\partial t^{2}}-\frac{\partial ^{2}\chi }{\partial x^{2}}+U_{sG}^{\prime \prime }\left( \Phi _{c}\right) \chi -2B\frac{\partial ^{2}\chi }{\partial x\partial t}+B^{2}\frac{\partial ^{2}\chi }{\partial x^{2}}=0 \label{51} \end{equation} We can now check, as promised, that constraint $C$ commutes with the quadratic Hamiltonian $H_{0}$: \begin{equation} \left[ C,H_{0}\right] =\int dx\Phi _{c}^{\prime }\left\{ \frac{\partial ^{2}\chi }{\partial x^{2}}-U_{sG}^{\prime \prime }\left( \Phi _{c}\right) \chi +B\frac{\partial \pi }{\partial x}\right\} +\frac{B}{\sqrt{1-B^{2}}}% \int dx\Phi _{c}^{\prime \prime }\left\{ \pi +B\frac{\partial \chi }{% \partial x}\right\} \label{51b} \end{equation} \noindent Integrate by parts so that the above expression becomes linear in $% \chi $ and $\pi $. The terms linear in $\pi $ cancel immediately. The bracket multiplying $\chi $ vanishes due to equation (\ref{4b}) obeyed by $% \Phi _{c}$. \section{Determination of the acceleration.} Constraint (\ref{420}) must commute with the total Hamiltonian $H$ (\ref{32}% ) and this determines $a$: \begin{equation} \left[ C,H_{c}\right] +B\left[ C,\int dx\Pi \frac{\partial \Phi }{\partial x}% \right] +a\left[ C,p_{B}\right] =0 \label{451a} \end{equation} \noindent An exact evaluation involves the fluctuating field in a somewhat cumbersome way. However in the first instance we are interested in the effect of the perturbing potential only (and disregard fluctuations). We use (\ref{35}), (\ref{36}), take commutators (\ref{37}), (\ref{38}) into account and keep terms proportional to $V$ only: \begin{equation} Ma\frac{d}{dB}\left( \frac{B}{\sqrt{1-B^{2}}}\right) =\int dx\Phi _{c}^{^{\prime }}\left( \frac{x}{\sqrt{1-B^{2}}}\right) V(x+X)U_{sG}^{^{\prime }}(\Phi _{c}) \label{451} \end{equation} \noindent or \begin{equation} M\frac{d}{dt}\left( \frac{B}{\sqrt{1-B^{2}}}\right) =\int dx\Phi _{c}^{^{\prime }}\left( \frac{x}{\sqrt{1-B^{2}}}\right) V(x+X)U_{sG}^{^{\prime }}(\Phi _{c}) \label{452} \end{equation} \noindent The left hand side is the relativistic expression for the force. The right hand side depends on $X$, hence (\ref{452}) cannot be readily integrated. The RHS of (\ref{452}) can be written \[ -\sin \left( k_{0}X\right) \int dx\Phi _{c}^{\prime }\left( \frac{x}{\sqrt{% 1-B^{2}}}\right) \Phi _{c}^{\prime \prime }\left( \frac{x}{\sqrt{1-B^{2}}}% \right) \sin \left( k_{0}x\right) \] \noindent where we used in turn (\ref{4b}) and (\ref{v}), expanded $V$ as a sum of cosines and sines and kept the sine part (the cosine part vanishing by parity). Rescaling \begin{equation} M\frac{d}{dt}\left( \frac{B}{\sqrt{1-B^{2}}}\right) =-\sqrt{1-B^{2}}\sin \left( k_{0}X\right) \int_{-\infty }^{\infty }dz\Phi _{c}^{\prime }(z)\Phi _{c}^{\prime \prime }(z)\sin \left( k_{0}z\sqrt{1-B^{2}}\right) \label{a2} \end{equation} \noindent Integrate the LHS\ by parts to get it in the form \[ \frac{1}{2}k_{0}\left( 1-B^{2}\right) \sin \left( k_{0}X\right) \int_{-\infty }^{\infty }dz\left( \Phi _{c}^{\prime }(z)\right) ^{2}\cos \left( k_{0}z\sqrt{1-B^{2}}\right) \] \noindent and using (\ref{zero}) \[ 2k_{0}\left( 1-B^{2}\right) \sin \left( k_{0}X\right) \int_{-\infty }^{\infty }dz\frac{\cos \left( k_{0}z\sqrt{1-B^{2}}\right) }{\cosh ^{2}z} \] \noindent The integral is standard and we get \begin{equation} M\frac{d}{dt}\left( \frac{B}{\sqrt{1-B^{2}}}\right) =\frac{2\pi k_{0}^{2}\left( 1-B^{2}\right) ^{3/2}}{\sinh \left( {\displaystyle {\pi k_{0}\sqrt{1-B^{2}} \over 2}} \right) }\sin \left( k_{0}X\right) \label{a3} \end{equation} The above expression connects relativistic acceleration and position. \section{The normal modes.} \subsection{The travelling modes.} \subsubsection{The (t,x) frame.} Solutions to (\ref{51}) are of the form \begin{equation} A(B,K,x)\exp (-i\Omega t+iKx) \label{52} \end{equation} \noindent \noindent and we set \[ g(B,K,x)=A(B,K,x)\exp (iKx) \] \noindent To determine the relation between $\Omega $ and $K$ we observe that away from the kink the modulating factor $A$ reduces to a constant; then (\ref{51}) yields \begin{equation} \Omega =-BK+\sqrt{K^{2}+\mu ^{2}} \label{60} \end{equation} \noindent Note that $K$ caries a sign depending on the direction of incidence. Also notice that $\Omega $ is positive (at least as long as the kink does not move at superluminal speeds). Relation (\ref{60}) will also be derived in the next paragraph by transforming from the inertial frame instantaneously moving with the kink. The explicit form of $A(B,K,x)$ can be deduced by struggling with (\ref{51}). We find it easier to deduce it in the next paragraph via a Lorentz transformation. It is clear from (\ref{51}) that due to the cross term in the derivatives the $g$s are not orthogonal. However we can still proceed and introduce creation and annihilation operators corresponding to the wavefunctions (\ref {52}), the procedure now being somewhat more complicated than in standard free field theory. We define (the index in $a^{\dagger }$ denoting the direction of incidence being suppressed) \begin{equation} a^{\dagger }(B,\Omega ,t)=\frac{1}{\sqrt{4\pi }}\int dxg^{*}(B,K,x)\left\{ \sqrt{\Omega }\chi -\frac{B}{i\sqrt{\Omega }}\frac{\partial \chi }{\partial x% }+\frac{\pi }{i\sqrt{\Omega }}\right\} \label{op1} \end{equation} \noindent with a corresponding expression for the complex conjugate. For $% B=0 $ this reduces to the usual expression. To check that $a^{\dagger }(B,\Omega ,t)$ satisfies the usual sinusoidal time evolution equation under the influence of the quadratic Hamiltonian $H_{0}$ rewrite (\ref{51}) in the form \begin{equation} (1-B^{2})\frac{d^{2}g}{dx^{2}}+U_{sG}^{\prime \prime }\left( \Phi _{c}\right) g+2iB\Omega \frac{dg}{dx}=-\Omega ^{2}g \label{op2} \end{equation} \noindent Commute the RHS\ of (\ref{op1}) with $H_{0}$ (\ref{50}), integrate by parts and use (\ref{op2}) to get in a straightforward manner \begin{equation} \frac{d}{dt}a^{\dagger }(B,\Omega ,t)=i\Omega a^{\dagger }(B,\Omega ,t) \label{op3} \end{equation} We shall later need the bracket $\left[ a^{\dagger }(B,\Omega ,t),\text{ }% p_{B}\right] .$ To this end use the definition (\ref{op1}) and (\ref{37}), (% \ref{38}) to get \begin{equation} \left[ a^{\dagger }(B,\Omega ,t),\text{ }p_{B}\right] =-\frac{i}{\sqrt{4\pi \Omega }}\frac{d}{dB}\left( \frac{B}{\sqrt{1-B^{2}}}\right) \displaystyle \int dxg^{*}(B,K,x)\Phi _{c}^{\prime }\left( \frac{x}{\sqrt{1-B^{2}}}\right) \label{op4a} \end{equation} \noindent with the complex conjugate for $a(B,\Omega ,t).$ Notice that in the nonrelativistic version the integral in the RHS of the above equation vanishes due to the orthogonality of the eigenfunctions of (\ref{5}). We expand the fluctuation field in the form \begin{equation} \chi (x,t)=\chi _{z}(x,t)+\frac{1}{\sqrt{2\pi }}\int \frac{d\Omega }{\sqrt{% \Omega }}\left\{ a^{\dagger }(B,\Omega ,t)g(B,K,x)+a(B,\Omega ,t)g^{*}(B,K,x)\right\} \label{op5} \end{equation} \noindent The above relation $defines$ the field $\chi _{z}(x,t)$ and is the relativistic analog of (\ref{16}) (with the attendant complications of non-orthogonality). To write the corresponding expansion for $\pi (x,t)$ we are inspired (i) by the relation \begin{equation} \frac{\partial \chi }{\partial t}=\pi +B\frac{\partial \chi }{\partial x} \label{op6} \end{equation} \noindent derived from $H,$ (ii) by the fact that the expansion should be valid in the special case of sinusoidal time evolution which holds when we neglect effects due to the potential or to terms of degree higher than quadratic in $\chi $. Thus \begin{eqnarray} \pi (x,t) &=&\pi _{z}(x,t)+\frac{i}{\sqrt{2\pi }}\int d\Omega \left\{ \sqrt{% \Omega }a^{\dagger }(B,\Omega ,t)g(B,K,x)-\sqrt{\Omega }a(B,\Omega ,t)g^{*}(B,K,x)\right\} + \nonumber \\ &&+\frac{1}{\sqrt{2\pi }}\int \frac{d\Omega }{\sqrt{\Omega }}\left\{ a^{\dagger }(B,\Omega ,t)\frac{\partial g(B,K,x)}{\partial x}+a(B,\Omega ,t)% \frac{\partial g^{*}(B,K,x)}{\partial x}\right\} \label{op7} \end{eqnarray} \noindent Again this is the definition of $\pi _{z}(x,t).$ We take the fields $\chi _{z}(x,t),\pi _{z}(x,t)$ to commute with the set of the $a$s and $a^{\dagger }$s. The bracket $\left[ \chi _{z}(x,t),\pi _{z}(x^{\prime },t)\right] $ is non-trivial in the nonrelativistic case as well. It equals \[ \frac{1}{M}\Phi _{c}^{\prime }(x)\Phi _{c}^{\prime }(x^{\prime }) \] \noindent as can be seen from the zero mode part of (\ref{16}), (\ref{17}). \subsubsection{The comoving frame.} In the absence of interactions the kink moves at uniform velocity. Denote quantities pertaining to the comoving frame by ($\sim $) and the lab frame coordinates by prime. Consider a travelling wave of the form (\ref{14}), (% \ref{14b}) \begin{equation} \exp \left( -i\tilde{\omega}\tilde{t}+i\tilde{p}\tilde{x}\right) \tilde{A}% \left( \tilde{p},\tilde{x}\right) \label{65} \end{equation} \noindent The phase appearing in (\ref{65}) is written in terms of primed coordinates \begin{equation} -\tilde{\omega}\frac{t^{\prime }-Bx^{\prime }}{\sqrt{1-B^{2}}}+\tilde{p}% \frac{x^{\prime }-Bt^{\prime }}{\sqrt{1-B^{2}}} \label{62a} \end{equation} \noindent Use (\ref{21}) for uniform $B$ and rearrange \begin{equation} -\tilde{\omega}\sqrt{1-B^{2}}t+\frac{\tilde{p}+B\tilde{\omega}}{\sqrt{1-B^{2}% }}x \label{62b} \end{equation} \noindent Compare with (\ref{52}) to identify \begin{equation} \Omega =\tilde{\omega}\sqrt{1-B^{2}}{\ , }K=\frac{\tilde{p}+B\tilde{\omega}}{% \sqrt{1-B^{2}}} \label{62c} \end{equation} \noindent Thus $(\Omega ,K)$ is a rather peculiar pair: $K$ coincides with the wavevector in the lab frame whereas $\Omega $ is connected to the frequency in the comoving frame. Manipulation of $\Omega ,K$ as given above yields again (\ref{60}). The Lorentz transformation together with (\ref{21}) yield the useful relation \begin{equation} \tilde{x}=\frac{x^{\prime }-Bt^{\prime }}{\sqrt{1-B^{2}}}\Rightarrow \tilde{x% }=\frac{x}{\sqrt{1-B^{2}}} \label{67} \end{equation} \noindent The same Lorentz transformation yields \begin{equation} \tilde{p}=\frac{K-B\sqrt{K^{2}+\mu ^{2}}}{\sqrt{1-B^{2}}} \label{66} \end{equation} \noindent (Recall that we our choice of units $\mu =1$.) Thus the mode (\ref {65}) is expressed in the $(t,x)$ frame in the form \begin{equation} \exp (-i\Omega t)g(B,K,x)=\exp (-i\Omega t+iKx)\tilde{A}\left( \tilde{p}(K),% \frac{x}{\sqrt{1-B^{2}}}\right) \label{66b} \end{equation} \noindent Hence (\ref{66b}) provides the explicit form of the solutions (\ref {52}) with $\tilde{A}$ taken from (\ref{14}), (\ref{14b}) and $\tilde{p}$ from (\ref{66}). In that sense it is hardly a mystery why operators $% a^{\dagger }(B,\Omega ,t)$, $a(B,\Omega ,t)$ diagonalize the total Hamiltonian: from the point of view of the comoving observer they correspond to the usual (orthogonal) modes (\ref{14}), (\ref{14b}) of frequency $\Omega /\sqrt{1-B^{2}}$. \subsection{The zero mode.} Consider an observer moving at velocity $B$ and suppose that the zero mode (% \ref{30}) is excited, i.e. \[ \chi (\tilde{x},\tilde{t})=u\tilde{t}\Phi _{c}^{\prime }\left( \tilde{x}% \right) \] \noindent where $u$ is a small parameter (the velocity of the kink with respect to the said observer). Transforming to the $\left( t,x\right) $ frame as above we get the same mode in the form \begin{equation} \chi (x,t)=ut\sqrt{1-B^{2}}\Phi _{c}^{\prime }\left( \frac{x}{\sqrt{1-B^{2}}}% \right) -u\frac{Bx}{\sqrt{1-B^{2}}}\Phi _{c}^{\prime }\left( \frac{x}{\sqrt{% 1-B^{2}}}\right) \label{70} \end{equation} \noindent It can be readily shown that the above expression satisfies (\ref {51}). This is precisely the form that $\chi _{z}(x,t)$ of (\ref{op5}) takes in the free case. From (\ref{op6}), (\ref{op7}) and (\ref{70}) we can work out the corresponding expression for $\pi _{z}(x,t).$ One can see constraint $C$ in a different light. It is easily shown that $C$ commutes with the $a$s and $a^{\dagger }$s and that in the special case of free fluctuations it turns out to be proportional to $u$ of (\ref{70}). In other words as seen from an observer moving at velocity $B$, the small velocity $u$ associated with the excitation of the zero mode vanishes if $C$ is to hold; this is precisely what one would expect from a kink moving at velocity $B$. \section{Radiation from an accelerated soliton.} To calculate the emission of radiation at a wavevector $K$ (with respect to the lab frame) we look at the commutator $\left[ a^{\dagger }(B,\Omega ,t),H\right] $ ($H$ being the total Hamiltonian). We write \begin{equation} \left[ a^{\dagger }(B,\Omega ,t),H\right] =(I)+(II)+(III) \label{71} \end{equation} $(I)$ is given by (\ref{op3}) (corresponding to free evolution). The commutator in (\ref{71}) gets contributions from terms in $H$ that are of order higher than quadratic in $\chi $. However we are only interested in contributions of first order in $V$ (i.e. of order $\varepsilon $; cf (\ref {v})). Contributions $(II)$ and $(III)$ come from commuting (\ref{op1}) with the term in the second pair of braces in (\ref{hkhi}) linear in $\chi $ and with the $p_{B}$ term respectively. \begin{eqnarray} (II) &=&-% {\displaystyle {\varepsilon \over i\sqrt{4\pi \Omega }}} \int_{-\infty }^{\infty }dxg^{*}(B,k,x)V(x+X)U_{sG}^{\prime }\left( \Phi _{c}\right) = \label{72} \\ &=&-% {\displaystyle {\varepsilon \over i\sqrt{4\pi \Omega }}} \int_{-\infty }^{\infty }dxg^{*}(B,k,x)\Phi _{c}^{\prime \prime }\left( \frac{x}{\sqrt{1-B^{2}}}\right) \cos \left( k_{0}x+k_{0}X\right) \nonumber \end{eqnarray} \noindent Recall that the $X$ has a non-trivial time dependence. For $(III)$ we get from (\ref{op4a}) \begin{equation} (III)=-a\frac{i}{\sqrt{4\pi \Omega }}\frac{d}{dB}\left( \frac{B}{\sqrt{% 1-B^{2}}}\right) \displaystyle \int dxg^{*}(B,K,x)\Phi _{c}^{\prime }\left( \frac{x}{\sqrt{1-B^{2}}}\right) \label{73} \end{equation} \noindent Given that $a=dB/dt$ the combination \[ a\frac{d}{dB}\left( \frac{B}{\sqrt{1-B^{2}}}\right) \] \noindent amounts to $1/M$ times the right hand side of (\ref{a3}). Notice that the RHS of (\ref{73}) does not vanish since the solutions of (\ref{351}% ) are not orthogonal. Thus \[ (III)=-\frac{i}{M\sqrt{4\pi \Omega }}\frac{2\pi k_{0}^{2}\left( 1-B^{2}\right) ^{3/2}}{\sinh \left( {\displaystyle {\pi k_{0}\sqrt{1-B^{2}} \over 2}} \right) }\sin \left( k_{0}X\right) \displaystyle \int dxg^{*}(B,K,x)\Phi _{c}^{\prime }\left( \frac{x}{\sqrt{1-B^{2}}}\right) \] \section{Conclusion.} This paper is meant to set the general framework for a treatment of the collective coordinate of a relativistic soliton interacting with fluctuations and an external potential. In a relativistic setting it proves natural to treat the velocity $B$ as a dynamical variable. This artificial increase in the number of degrees of freedom and the attendant gauge invariance generated by the constraint (\ref{27}) are lifted by the subsidiary condition that defines the position of the soliton. Let us stress again that the formulae for the total momentum and energy of the system (\ref {41}) and (\ref{400c}) include the correct relativistic expressions for the soliton momentum and energy. It is equally reassuring that the left hand sides of (\ref{452}) and (\ref{a3}) feature the time derivative of the relativistic momentum. Concerning emission of radiation we observe that it has two physical origins. One is the interaction of the soliton with the fluctuations. This comes about when we commute the $\pi $ dependent term in (\ref{op1}) with the $V$ independent cubic and quartic terms in $\chi $ in the total Hamiltonian. This contribution was disregarded in the present work only because we were interested in the $V$ dependence. The acceleration also receives contributions from terms in the Hamiltonian depending on both $V$ and $\chi ,$ reflecting the fact that an accelerated soliton radiates; to lowest order we obtain ($II)$ calculated above. Finally observe that contribution $(III)$ stems from the $p_{B}$ term in the Hamiltonian, whose existence is required by the need of a relativistic treatment. \section{Acknowledgments.} The author is indebted to Professor Gabriel Barton for his comments and to Professor Grigori Volovik for discussions. He also wishes to thank the Low Temperature Laboratory of Helsinki University of Technology for its hospitality and EU\ TMR programme ERBFMGECT980122 for support.
\section{Introduction\label{sec:introduction}} \nopagebreak The importance of bars in the structure of spiral galaxies is recognised in the Hubble sequence for the classification of galaxies (Sandage \markcite{s61}1961). In the first paper of this series (Bureau \& Athanassoula \markcite{ba99}1999, hereafter Paper~I), we highlighted the difficulties involved in the identification of bars in edge-on systems. It is clear that the detection of such bars based on photometric or morphological criteria (e.g.\ de Carvalho \& da Costa \markcite{dd87}1987; Hamabe \& Wakamatsu \markcite{hw89}1989) is uncertain. Kuijken \& Merrifield (\markcite{km95}1995, hereafter KM95; see also Merrifield \markcite{m96}1996) were the first to show that the particular kinematics of barred disks could be used to identify bars in edge-on spirals. They showed that the periodic orbits in an edge-on barred galaxy produce characteristic double-peaked line-of-sight velocity distributions, which can be taken as the signature of a bar. In \markcite{ba99}Paper~I, we improved on the work of \markcite{km95}KM95. We studied the signatures of individual periodic orbit {\em families} in the position-velocity diagrams (PVDs) of edge-on barred spirals (before combining them to model real galaxies), and examined how the PVDs depend on the viewing angle. We adopted a widely used mass model and a well-defined method to populate the periodic orbits. Our aim was to provide insight into the projected kinematical structure of barred disks, and to provide guidance to interpret stellar and gaseous kinematical observations of edge-on spiral galaxies. We showed in \markcite{ba99}Paper~I that the global appearance of a PVD can be used as a diagnostic to detect the presence of a bar in an edge-on disk. The signatures of the various periodic orbit families leave gaps in the PVDs which are a direct consequence of the non-homogeneous distribution of orbits in a barred spiral. The signature of the $x_1$ periodic orbits is parallelogram-shaped and occupies all four quadrants of the PVDs, reaching very high radial velocities when the bar is seen end-on and only low velocities when the bar is seen side-on. The signature of the $x_2$ orbits, when present, is similar to that of the $x_1$, but reaches its maximum radial velocities at opposite orientations. Those features can be used to determine the viewing angle with respect to the bar in an edge-on disk. However, even if carefully chosen and populated, periodic orbits provide only an approximation to the structure and kinematics of the stars and gas in spiral galaxies. For example, a number of stars may be on irregular orbits, and shocks can develop in the gas. In this paper (Paper~II), we concentrate on developing bar diagnostics for edge-on disks using the gaseous component alone. We use the hydrodynamical simulations of Athanassoula (\markcite{a92b}1992b, hereafter \markcite{a92b}A92b), designed to study the gas flow and shock formation in barred spiral galaxies. Unlike \markcite{ba99}Paper~I, these simulations properly take into account the fact that the gas is not a collisionless medium. The shocks and inflows which develop in the simulations lead to better bar diagnostics than those of \markcite{ba99}Paper~I. In addition, we run simulations covering a large fraction of the parameter space likely to be occupied by real galaxies. The PVDs produced can thus be directly compared with observations not only to detect the presence of bars in edge-on spiral galaxies, but also to constrain the mass distribution of the systems observed. In particular, Bureau \& Freeman (\markcite{bf99}1999, hereafter BF99) have applied those diagnostics to long-slit spectroscopic observations of a large number of edge-on spiral galaxies, most of which have a boxy or peanut-shaped bulge, to determine the formation mechanism of these objects and study the vertical structure of bars. In \markcite{ab99}Paper~III (Athanassoula \& Bureau \markcite{ab99}1999), using fully self-consistent three-dimensional (3D) $N$-body simulations, we will develop similar bar diagnostics for the stellar (collisionless) component of barred spiral galaxies. We describe the mass model and hydrodynamical simulations used in this paper in \S~\ref{sec:hydro}. In \S~\ref{sec:bardiag}, we study the signatures in the PVDs of the various components present in the simulations and discuss the influence of the parameters of the mass model. The effects of dust extinction are illustrated in \S~\ref{sec:dust}. We develop the bar diagnostics for edge-on disks and discuss the limitations of our models for the interpretation of real data in \S~\ref{sec:discussion}. We conclude in \S~\ref{sec:conclusions} with a brief summary of our main results. \section{Hydrodynamical Simulations\label{sec:hydro}} \nopagebreak For the hydrodynamical simulations, we use the flux-splitting second-order scheme of G.\ D.\ van Albada (van Albada \& Roberts \markcite{vr81}1981; van Albada, van Leer, \& Roberts \markcite{vvr82}1982; van Albada \markcite{v85}1985). It is the same code as that used by \markcite{a92b}A92b so we will only briefly review its main properties here. The simulations are two-dimensional and time-dependent, and the gas is treated as ideal, isothermal, and non-viscous. The simulations are not self-consistent and we do not consider the self-gravity of the gas; the flow is calculated using the potential described in \markcite{ba99}Paper~I (see also Athanassoula \markcite{a92a}1992a, hereafter \markcite{a92a}A92a). The mass model has two axisymmetric components, a Kuzmin/Toomre disk (Kuzmin \markcite{k56}1956; Toomre \markcite{t63}1963) and a bulge-like spherical density distribution. They combine to yield a flat rotation curve in the outer parts of the disk. We use a homogeneous ($n=0$) or inhomogeneous ($n=1$) Ferrers spheroid (Ferrers \markcite{f77}1877) as a third component representing the bar. Each model is described by four main parameters: the bar axial ratio $a/b$, the quadrupole moment of the bar $Q_m$ (proportional to the mass of the bar), the Lagrangian radius $r_L$ (the radius of the Lagrange points $L_1$ and $L_2$ on the major axis of the bar, approximately inversely proportional to the bar pattern speed), and the central concentration $\rho_c$. The other quantities are fixed, including the semi-major axis of the bar $a=5$~kpc. We refer the reader to \markcite{ba99}Paper~I for a more detailed description of the mass model. \markcite{a92a}A92a discusses at great length its relevance to real galaxies. Suffice it to say here that the properties of the mass model are in excellent agreement with those of early-type barred spirals. The simulations are started with a massless bar and an additional axisymmetric component of mass equal to that of the desired final bar. Mass is transfered from that component to the bar over $10^8$~yr and the simulations are run until the gas flow is roughly stationary in the frame of reference corotating with the bar (about 10 bar revolutions). An $80\times160$~cells grid is used to cover a disk of 16~kpc radius, assuming bisymmetry. The inner half of the simulations are then regridded to the $80\times160$ grid and the simulations continued for another 5 bar revolutions using this increased spatial resolution ($0.1\times0.1$~kpc$^2$ cell size; see van Albada \& Roberts \markcite{vr81}1981). Star formation and mass loss are modeled crudely. The gas density is lowered artificially in high density regions and gas is added uniformly over the grid. This process is governed by the equation \begin{equation} \label{eq:sf} d\rho_g/dt=\alpha\rho_{g,init}^2-\alpha\rho_g^2, \end{equation} where $\rho_g$ is the gas density, $\alpha$ is a constant (set to 0.3~$M_{\mbox{\scriptsize \sun}}$ $^{-1}$~pc$^2$~Gyr$^{-1}$ in most runs), and $\rho_{g,init}=1$~$M_{\mbox{\scriptsize \sun}}$ ~pc$^{-2}$ is the uniform initial gas density. It therefore takes about 3~Gyr for the gas to be reprocessed. There is no artificial viscosity in the code. Our main tool in this paper will be PVDs, representing the projected density of material in the edge-on disks as a function of line-of-sight velocity and projected position along the major axis. Since we are only interested in the bar region, we use only the inner 8~kpc~$\times$~8~kpc region of the simulations, covered by the high resolution grid. At larger radii, the motion of the gas can, to first order, be considered as circular, since the force of the bar decreases steeply with radius. Including the outer regions thus makes no difference to our PVDs. As in \markcite{ba99}Paper~I, the models considered are those of \markcite{a92a}A92a (see her Table~1). We will also use her units: $10^6$~$M_{\mbox{\scriptsize \sun}}$ for masses, kpc for lengths, and km~s$^{-1}$ for velocities. It is essential to understand the orbital structure of the models to interpret properly the results of the simulations. Orbital properties have been discussed in \markcite{ba99}Paper~I and \markcite{a92a}A92a, but also in Athanassoula et al.\ \markcite{abmp83}(1983), Papayannopoulos \& Petrou \markcite{pp83}(1983), Teuben \& Sanders \markcite{ts85}(1985), and others. Here, we will mainly use \markcite{ba99}Paper~I and \markcite{a92a}A92a for comparison. We will also draw heavily on the results of \markcite{a92b}A92b, which used the same set of simulations but for different purposes. She discussed in detail the gas flow and compared the results with the properties of real galaxies. \section{Bar Diagnostics\label{sec:bardiag}} \nopagebreak In this section, we will concentrate on understanding the PVDs of two inhomogeneous bar models which are prototypes of models with and without inner Lindblad resonances (ILRs). As suggested in \markcite{a92a}A92a, we will identify the existence and position of the ILRs with the existence and extent of the $x_2$ periodic orbits. The two models considered are the same as in \markcite{ba99}Paper~I: model 001 ($n=1$, $a/b=2.5$, $Q_m=4.5\times10^4$, $r_L=6.0$, $\rho_c=2.4\times10^4$) and model 086 ($n=1$, $a/b=5.0$, $Q_m=4.5\times10^4$, $r_L=6.0$, $\rho_c=2.4\times10^4$). We will then extend our results to other models and analyse how the properties of the PVDs vary as the parameters of the mass model are changed. \subsection{Model 001 (ILRs)\label{sec:001}} \nopagebreak \placefigure{fig:001} Figure~\ref{fig:001} shows PVDs for model 001, which has ILRs. The figure shows, for the inner half of the simulation, the face-on surface density of the gas and PVDs obtained by projecting the simulation edge-on and using various viewing angles with respect to the bar. Unlike the models of \markcite{ba99}Paper~I, which are symmetric around both the minor and major axes of the bar, the present simulations are bisymmetric, so we need to cover a viewing angle range of 180\arcdeg. The viewing angle $\psi$ is defined to be 0\arcdeg\ for a line-of-sight parallel to the major axis of the bar and 90\arcdeg\ for a line-of-sight perpendicular to it, increasing counterclockwise in the surface density plots. \markcite{a92b}A92b showed convincingly that the two strong parallel narrow segments present in the surface density plot of Figure~\ref{fig:001} represent offset shocks on the leading sides of the bar, displaying strong density enhancements and sharp velocity gradients. They can be identified with the dust lanes observed in barred spiral galaxies. The structures seen at the ends of the bar and perpendicular to it are also shocks. In the inner bar region is a very intense two-arm nuclear spiral, connecting with the offset shocks. There is little gas in the barred region outside the nuclear spiral, which we will hereafter refer to as the outer bar region. Beyond the bar the surface density is almost featureless; only a few spiral arms are seen. \markcite{a92b}A92b showed that, in the outer bar region, the streamlines have roughly the shape and orientation of the $x_1$ periodic orbits. As one moves inward, the streamlines change gradually to the shape and orientation corresponding to the $x_2$ periodic orbits. This flow pattern leads to the offset shocks and results in an inflow of gas toward the nuclear region, accounting for the gas distribution in the bar: low densities in the outer bar region and high densities in the center. The velocities are small (in the reference frame corotating with the bar) around the Lagrange points on the minor axis of the bar and the flow is close to circular outside the bar region. The PVDs in Figure~\ref{fig:001} show the existence of three distinct regions: an inner region, corresponding to the signature of the nuclear spiral; an intermediate region, corresponding to the signature of the outer parts of the bar; and an external region, corresponding to the signature of the parts outside the bar. It is important to notice how low the density of the intermediate region is, compared to that of the other two regions. This is not surprising since, as mentioned in \markcite{a92b}A92b and above, most of the bar region has very low gas density, with the exception of the central part (the nuclear spiral) and the two shock loci. This will form the basis of the bar diagnostics which we will develop later on. Let us now examine each of the three regions separately, by keeping only the gas in the targeted region (and masking out the gas in the other regions) when calculating the PVDs. \placefigure{fig:bar} Figure~\ref{fig:bar} shows the surface density and PVDs of model 001 when considering the low density outer bar region only. Because the high density regions have been masked out, and because we are looking only at relatively low density regions, we see much more structure in the PVDs of Figure~\ref{fig:bar} than in the corresponding sections of the PVDs of Figure~\ref{fig:001}. When compared with Figure~4 of \markcite{ba99}Paper~I, Figure~\ref{fig:bar} shows, at first glance, many similarities, but also, when scrutinised closer, a number of differences. This could be expected since the gas streamlines in that region follow loosely, but far from exactly, the shape and orientation of the $x_1$ orbits, which are elongated parallel to the bar. The parallelogram-shaped signature of the $x_1$ orbits in Figure~4 of \markcite{ba99}Paper~I is again observed here and the ``forbidden'' quadrants are again populated. This is due to the fact that both the $x_1$ orbits and the streamlines in the outer bar region are not circular, but elongated. In the gas, however, the parallelogram shape is also present for $\psi=0\arcdeg$, while, for this viewing angle, the PVD of the $x_1$ orbits showed a bow-shaped feature. The reason is that, unlike the $x_1$ orbits, the streamlines do not have their major axes parallel to that of the bar, but rather at an angle of about 20\arcdeg\ to it (see \markcite{a92b}A92b). Also, at $\psi=90\arcdeg$, the $x_1$ orbits show a near-linear PVD, while the gas shows an additional faint bow-shaped feature reaching high velocities near the center. Masking out a larger fraction of the simulation (not shown), it is possible to show that this extra feature arises from the very low density areas just inside the offset shocks, close to the major axis of the bar. In this section of the bar, the flow differs substantially from the behaviour of the $x_1$ periodic orbits. The elongated ``hole'' in the center of the PVDs, at intermediate viewing angles, is present in both sets of PVDs, and is due to the way the $x_1$ family was terminated and the way the gas density was masked out. This causes in both cases an absence (or low density) of quasi-circular streamlines near the end of the bar. In the hydrodynamical simulation, there is also a reduced degree of symmetry with respect to viewing angles of 0\arcdeg\ or 90\arcdeg, in particular for the angles 67.5\arcdeg\ and 112.5\arcdeg. They are, of course, identical in the PVDs of \markcite{ba99}Paper~I. This is because the gas flow is bisymmetric, as opposed to being symmetric with respect to both the minor and major axes of the bar, as are the $x_1$ orbits (see Fig.~2 of \markcite{a92b}A92b). In Figure~\ref{fig:bar}, we see that the radial velocities are maximum (compared to the circular velocity) for lines-of-sight almost parallel to the bar (highest values for $\psi=0\arcdeg$ and 157.5\arcdeg), and decrease as $\psi$ increases, to reach a minimum when the line-of-sight is roughly along the bar minor axis (lowest values for $\psi=90\arcdeg$ and 67.5\arcdeg). The same behaviour was seen for the PVDs of the $x_1$ orbits (\markcite{ba99}Paper~I), but with a slight shift of the viewing angle, so that the maximum and minimum occurred at exactly 0\arcdeg\ and 90\arcdeg, respectively. This was to be expected since, like the $x_1$ orbits, the streamlines in the outer bar region are elongated parallel to it (with a small offset). Similarly, the position where the maximum occurs moves out as the viewing angle increases from $\psi=0\arcdeg$ to $\psi=90\arcdeg$. This was also found for the $x_1$ orbits (\markcite{ba99}Paper~I), where it was explained by considering the trace of an elongated orbit in a PVD as a function of viewing angle and distance from the center. The explanation is analogous here. However, in the hydrodynamical simulation, the maxima generally occur at a larger distance from the center than in the periodic orbits approach (this is easily seen by comparing Fig.~\ref{fig:bar} to Fig.~4 in \markcite{ba99}Paper~I). This is due to the fact that the $x_1$-like flow does not extend up to the center of the simulated galaxy (contrary to the $x_1$ orbits in \markcite{ba99}Paper~I), but is superseded by an $x_2$-like flow at a certain radius. For example, the nuclear spiral (which has an $x_2$-like behaviour) extends about 1.6~kpc on the minor axis of the bar, which is the projected distance at which the envelope of the signature of the outer bar region in the PVD for $\psi=0\arcdeg$ drops abruptly. To see which features of the gas distribution contribute to the PVDs, we have repeated the blanking exercise, this time masking out all areas except either the shock loci along the leading sides of the bar or the density enhancements near the ends of the bar major axis (not shown). The gas in the shock loci does not contribute any outstanding features to the PVDs, mainly because the amount of gas integrated along most lines-of-sight is small, the features being very narrow. At $\psi=22.5\arcdeg$, when the line-of-sight is nearly parallel to the shock loci, they contribute the two straight and parallel segments separating the intense and faint regions of the PVD just inside ${\mbox D}=\pm 2$. The high density enhancements near the ends of the bar major axis contribute the very strong linear segment going through the center of the PVDs at $\psi=0\arcdeg$ and 22.5\arcdeg. For these angles, they are seen roughly as segments of circles. For viewing angles near 90\arcdeg, they give rise to the undulating parts of the PVDs at large radius. \placefigure{fig:nuclear} We can repeat the masking process to keep only the region of the simulation containing the nuclear spiral, as shown in Figure~\ref{fig:nuclear}. Comparing the PVDs thus obtained with those of the $x_2$ family of periodic orbits (Fig.~5 of \markcite{ba99}Paper~I), we find again many similarities, but also some differences. In both cases, we see either an inverted S-shaped feature or a near straight segment passing through the center of the PVDs. We refer the reader to \markcite{ba99}Paper~I for a discussion and explanation of these shapes. More similarity is found, however, if we compare PVDs at viewing angles differing by $\approx20\arcdeg$, e.g.\ comparing the gaseous PVD at $\psi=112.5\arcdeg$ to the orbital one at $\psi=90\arcdeg$. This offset seems to corresponds to an offset of the nuclear spiral with respect to the minor axis of the bar (and to the offset of the straight shocks with respect to the major axis, see \markcite{a92b}A92b), although this is hard to measure in the surface density plot. Also, the PVDs of the nuclear spiral are more asymmetric with respect to the viewing angles 0\arcdeg\ or 90\arcdeg\ than the PVDs of the outer bar region (e.g.\ comparing the PVDs at $\psi=22.5\arcdeg$ and $\psi=-22.5\arcdeg=157.5\arcdeg$). This is due to the weaker symmetry of the nuclear spiral. Because the streamlines in the central region are elongated roughly perpendicular to the bar, like the $x_2$ orbits, higher radial velocities (compared to the circular velocity) are reached when the bar is seen side-on than when the bar is seen end-on. Considering the maximum velocity of the PVDs as a function of the viewing angle, we find that it is highest for $\psi=112.5\arcdeg$ and 135\arcdeg\ and lowest for $\psi=0\arcdeg$ and 22.5\arcdeg. The ``hole'' in the parallelogram-shaped signature of the $x_2$ orbits (see Fig.~5 of \markcite{ba99}Paper~I) has completely disappeared in the hydrodynamical simulation, because the $x_2$-like behaviour of the gas flow persists past the inner ILR and the flow becomes almost circular in the very center. \placefigure{fig:outer} Figure~\ref{fig:outer} isolates the signature of the outer parts of the simulation in the PVDs. Because the influence of the bar decreases rapidly with radius, the flow outside the bar is close to circular. A perfectly circular orbit would yield an identical inclined straight line passing through the origin in all PVDs. The structure seen here can thus be thought of as a succession of near-circular orbits of increasing radii, yielding the ``bow tie'' signature observed. The ``hole'' seen in the center of the PVDs at certain viewing angles (e.g.\ $\psi=45\arcdeg$) is again due to the fact that the orbits are not perfectly circular. The strong almost solid-body features forming loop-like structures near the upper and lower limits of the envelope of the signature are due to tightly wound spiral arms in the outer disk. \subsection{Model 086 (no-ILRs)\label{sec:086}} \nopagebreak \placefigure{fig:086} Figure~\ref{fig:086} shows the face-on surface density distribution of model 086, which has no ILRs (or, equivalently, has no $x_2$ periodic orbits), and the PVDs obtained using an edge-on projection. The main difference with the density distribution of model 001 is the absence of any significant nuclear spiral. The strong straight and narrow features in the center of the bar are centered shock loci, caused by the high curvature of the streamlines near the major axis of the bar (see Fig.~4 in \markcite{a92b}A92b). Similarly, the $x_1$ orbits in this model have high curvatures or loops near their major axes (\markcite{a92a}A92a). The shock loci do not curve near the center because no streamline perpendicular to the bar exists; there are no $x_2$ periodic orbits in this model. The shocks, however, still drive an inflow of gas, resulting in the {\em entire} bar region being gas depleted. As expected, the region outside the bar is almost unaffected by the change in the bar axial ratio compared to model 001. The PVDs for model 086 reflect all of the above properties. In particular, they lack the central feature associated with the nuclear spiral in model 001, and the signature of the bar region is very faint (Fig.~\ref{fig:086} should be contrasted with Fig.~\ref{fig:001}). Because of the absence of an $x_2$-like flow in the center of the simulation, the $x_1$-like flow extends all the way to the center. In good agreement with what happens for the periodic orbits (see \markcite{ba99}Paper~I or \markcite{a92a}A92a), the streamlines are much more eccentric than those of model 001. The parallelogram-shaped envelope of the signature of the bar region in the PVDs therefore reaches even more extreme radial velocities (compared to the circular velocity) than in model 001, and the rising part of the envelope extends almost all the way to the center for viewing angles close to 0\arcdeg\ (see also Fig.~9 in \markcite{ba99}Paper~I). When the bar is seen close to end-on, radial velocities more than twice those in the outer parts are reached. These velocities decrease rapidly as the viewing angle approaches $\psi=90\arcdeg$. Unfortunately, the signature of the bar region in the PVDs is very faint, because of the strong inflow of gas toward the center. As in the case of model 001, the regions outside the bar lead to a strong almost solid-body component in all PVDs. \subsection{Other Models\label{sec:others}} \nopagebreak In this section, we will analyse sequences of simulations corresponding to the ranges along the axes of parameter space likely to be occupied by real galaxies. We hope thereby to further our understanding of the PVDs of edge-on barred spiral galaxies, and to be able to extend the bar diagnostics which we will develop. We also wish to develop criteria allowing us to constrain the bar properties and mass distribution of edge-on systems. To achieve this, we will concentrate on understanding the PVDs of the gas flow within the barred region of the simulations, and will use extensively for that purpose the results of \markcite{a92a}A92a and \markcite{a92b}A92b. In particular, \markcite{a92b}A92b showed that nuclear spirals occur in models with an extended $x_2$ family of periodic orbits and lead to offset shocks, while centered shocks occur in models with no or shortly extended $x_2$ orbits. She also showed that $x_1$ periodic orbits with a high curvature near the major axis of the bar are essential to the formation of shocks, and that such shocks lead to an inflow of gas toward the central regions of the simulations, depleting the outer (or entire) bar regions. \placefigure{fig:a/b} Figure~\ref{fig:a/b} shows a sequence of simulations with varying bar axial ratio $a/b$ (the other parameters are kept fixed at those of model 001). The figure shows, for each simulation, the face-on surface density distribution and the velocity field (with a few streamlines) in the frame of reference corotating with the bar. It also shows the PVDs obtained for various viewing angles with respect to the bar when the simulation is viewed edge-on. To limit the size of the figure, we show a slightly reduced number of viewing angles compared to Figures~\ref{fig:001}--\ref{fig:086}. For small bar axial ratios, when the shocks are short and curved, the outer bar region is not strongly gas depleted and its signature is easily visible in the PVDs. As we consider simulations with increasingly high bar axial ratios, the density of the region of the PVDs which corresponds to the outer bar region drops considerably. It reaches its lowest values for the highest bar axial ratio considered ($a/b=5.0$), for which the shock loci are straight and close to the bar major axis. The envelope of the signature of the outer bar region in the PVDs becomes more extreme as the bar axial ratio is increased: the maximum radial velocities reached for small viewing angles increase and the positions of the maxima get closer to the center. These behaviours are quasi-linear and the opposite is true for large viewing angles. For example, at $\psi=0\arcdeg$, the maximum velocity of the outer bar region is about 250 for $a/b=1.5$, 360 for $a/b=3.0$, and 430 for $a/b=5.0$. At $\psi=90\arcdeg$, the opposite is observed, with maximum velocities of about 225 and 215, respectively. This is easily understood because, as the bar axial ratio increases, the eccentricity of the streamlines in the outer bar region (and that of the $x_1$ orbits, see \markcite{a92a}A92a) also increases, leading to higher velocities along the line-of-sight when the bar is seen end-on, and lower velocities when the bar is seen side-on. The most important effect of an increase of the bar axial ratio is the disappearance of the nuclear spiral for axial ratios $a/b\gtrsim2.7$. Because the nuclear spiral is associated with an $x_2$-like flow, and because the range of radii occupied by the $x_2$ orbits decreases rapidly as the bar axial ratio is increased (see Fig.~6 of \markcite{a92a}A92a), the inverted S-shaped signature of the nuclear spiral in the PVDs disappears for large bar axial ratios. The maximum radial velocity reached by the nuclear spiral signature varies little with the bar axial ratio (when present). \placefigure{fig:rl} Figure~\ref{fig:rl} shows how the PVDs of the simulated disks change when the Lagrangian radius of the mass model is varied. \markcite{a92b}A92b showed that the gas flow in a bar has shock loci offset toward the leading sides of the bar and of the form observed in early-type barred spiral galaxies only for a restricted range of Lagrangian radii, namely $r_L=(1.2\pm0.2)a$. All the observational estimates for early-type strongly barred spirals also give values within this range (see, e.g., \markcite{a92b}A92b; Elmegreen \markcite{e96}1996). We will thus concentrate on this range here. Model 028, the limiting case with $r_L=5.0$, shows strong spiral arms starting at the ends of the bar and extending to large radii. The spiral arms are easily identified in the PVDs as long filamentary structures. This model has no nuclear spiral, and thus no corresponding inverted S-shaped feature in the PVDs. The opposite is true for models with larger Lagrangian radii or, equivalently, lower pattern speeds. Those models have extended $x_2$ families of periodic orbits and therefore nuclear spirals (see \markcite{a92b}A92b). As the pattern speed of the bar is further decreased, the radial range occupied by the $x_2$ periodic orbits is increased and the nuclear spiral becomes more predominant. The outer bar region decreases accordingly (\markcite{a92a}A92a). This effect is clearly seen in the PVDs of Figure~\ref{fig:rl}. For increasing Lagrangian radii, they show an increase of the radial extent of the nuclear spiral signature, and a decrease of the radial extent of the outer bar region signature. There is also an increase of the maximum radial velocity reached by the nuclear spiral, but the effect is rather small. \placefigure{fig:rhoc} Figure~\ref{fig:rhoc} shows the behaviour of the PVDs as the central concentration of the mass model is varied. For low central concentrations, the gas streamlines are oval-shaped and aligned with the bar, there are no shocks, and the bar region is only slightly gas depleted. The signature of the bar region in the PVDs is then clear. As the central concentration is increased, the streamlines become more eccentric and the envelope of the signature of the bar region extends to higher velocities. Once the central concentration reaches $\rho_c\approx2.2\times10^4$, an $x_2$-like flow appears in the center of the bar and a nuclear spiral and offset shocks are formed. These changes are also easily seen in the PVDs, which acquire an inverted S-shaped feature, while the density in the outer bar region drops. The region occupied by the $x_2$ orbits then increases with increasing central concentration (\markcite{a92a}A92a), and so does the radial extent of the nuclear spiral signature in the PVDs. An increase in the central concentration of the mass model has similar effects to an increase of the Lagrangian radius (see Fig.~\ref{fig:rl}). This is not surprising, since both changes influence the location of the resonances in a similar way. \placefigure{fig:qm} Figure~\ref{fig:qm} shows a sequence of simulations with varying bar quadrupole moment. For the lowest quadrupole moment, the bar is weak and the flow is close to circular, with only a weak nuclear spiral and curved shocks in the center. This does not cause substantial inflow, and the gas density in the barred region remains high. The velocity field shows that there is a transition from a $x_1$-like flow to a $x_2$-like flow in the central region, but the effect is not strong since the eccentricity of the streamlines is small. All these effects reflect themselves in the PVDs. For somewhat higher bar quadrupole moments (model 058), the nuclear spiral is well-developed and its signature is strong in the PVDs, with a gap present between it and the solid-body signature of the outer parts of the galaxy. As the bar quadrupole moment is increased further, the envelope of the signature of the outer bar region in the PVDs becomes more extreme, reaching larger radial velocities and extending closer to the center. For $Q_m\gtrsim5.5\times10^4$, the nuclear spiral disappears. Those two effects are due to the facts that the eccentricity of the streamlines increases significantly with increasing bar quadrupole moment, while the region occupied by the $x_2$ orbits decreases until it disappears completely (see \markcite{a92a}A92a). Because the bar is so strong for high quadrupole moments, the flow is non-circular even in the outer parts of the simulations, and a ``hole'' is present in the center of the PVDs at intermediate viewing angles. Not surprisingly, the variations in the gas distribution and kinematics are similar when the bar quadrupole moment is increased and when the axial ratio of the bar is increased. \section{Dust Extinction\label{sec:dust}} \nopagebreak It is interesting to note that the PVDs discussed so far have a certain degree of symmetry with respect to the viewing angles 0\arcdeg\ or 90\arcdeg. This means that although it is relatively easy to determine whether a line-of-sight is close to the major or the minor axis of a bar ($|\psi|$), it is considerably harder to determine in which half of the PVD (positive or negative projected distances from the center) the near side of the bar is located ($\pm\psi$). This situation can be contrasted to that in the Galaxy, where most studies have no difficulty identifying the quadrant in which the near side of the Galactic bar is located. Studies using infrared photometry (e.g.\ Dwek et al.\ \markcite{detal95}1995; Binney, Gerhard, \& Spergel \markcite{bgs97}1997), star counts (e.g.\ Weinberg \markcite{w92}1992; Stanek et al.\ \markcite{suskkmk97}1997), gaseous or stellar kinematics (e.g.\ Binney et al.\ \markcite{bgsbu91}1991; Wada et al.\ \markcite{wthh94}1994; Zhao, Spergel, \& Rich \markcite{zsr94}1994; see also Beaulieu \markcite{b96}1996; Sevenster et al.\ \markcite{ssvf97}1997), and microlensing events (e.g.\ Paczynski et al.\ \markcite{psuskkmk94}1994) all indicate a bar making an angle of 15\arcdeg\ to 45\arcdeg\ with respect to the line-of-sight to the Galactic center (positive values indicating that the near side of the bar is at positive Galactic longitude). In the case of the Galaxy, at least two effects help the observer determine the exact orientation of the bar. Firstly, projection effects (both in longitude and latitude) mean two lines-of-sight on each side of the Galactic center reach correspondingly different parts of the bar (e.g.\ Binney et al.\ \markcite{bgsbu91}1991). Secondly, the large difference between the distances to each side of the bar means point sources in the far side of the bar will appear significantly fainter than the corresponding sources in the near side (e.g.\ Stanek et al.\ \markcite{suskkmk97}1997). In addition, the far side of the bar will appear thinner that the near side (Dwek et al.\ \markcite{detal95}1995). For a galaxy at infinity, all lines-of-sight are parallel, and no projection or distance effects are present. However, extinction within an edge-on disk can play a similar role to that of distance in the Galaxy, and can help constrain the orientation of a bar. Because the velocity spread in the inner parts of the PVDs is so large, it is unlikely that self-absorption by any line would be significant. Given the prominence of the dust lane in many edge-on spiral galaxies, extinction by dust is likely to be the dominant factor affecting the PVDs. If dust is present in significant amount, the spectroscopic signature of the far side of the bar in a PVD should be fainter than that of the nearer side. \placefigure{fig:dust} Figure~\ref{fig:dust} shows the surface density and PVDs of model 001 when considering a dust distribution proportional to the gas surface density. Because our simulations are not self-consistent, only the relative values of the density are important and the value of the dust absorption coefficient per unit mass $\kappa$ we use is meaningless. Our goal in this section being merely to illustrate the effects of dust on the simulated PVDs, and not to reproduce quantitatively the situation in real galaxies, we have simply increased the value of $\kappa$ (in which the dust to gas ratio is also folded) until the PVDs were significantly affected. In Figure~\ref{fig:dust}, we have decreased the surface densities in the surface density plot to reflect the effective contribution of each point to the projected density for a viewing angle $\psi=45\arcdeg$. Unlike Figure~\ref{fig:001}, the PVDs are now far from symmetric with respect to viewing angles of 0\arcdeg\ or 90\arcdeg. In addition, PVDs at intermediate viewing angles are no longer antisymmetric with respect to the center. This is mainly because the nuclear spiral obscures most of the material behind it, breaking the symmetry of the parallelogram-shaped signature of the $x_1$-like flow in the outer bar region. For viewing angles $0\arcdeg\lesssim\psi\lesssim90\arcdeg$, the nuclear spiral obscures mostly material moving away from the observer in the outer bar region, and obscures only a small amount of material moving toward the observer in the same region. Thus, the signature of the $x_1$-like flow in the PVDs is weakened for positive radial velocities, leading to a much fainter signature of the outer bar region in the upper halves of the PVDs than in the lower halves. The opposite is true for viewing angles $90\arcdeg\lesssim\psi\lesssim180\arcdeg$, where the signature of the outer bar region is much fainter in the lower halves. This effect is strongest and least extended for viewing angles $\psi\approx110\arcdeg$ to 135\arcdeg, as the line-of-sight is then roughly parallel to the major axis of the nuclear spiral (which is slightly offset from that of the $x_2$ periodic orbits). Some effects on the signature of the nuclear spiral itself and on the signature of the outer parts of the simulation are present, but they are less pronounced. In addition to the diagnostics suggested in the previous sections to identify a bar in an edge-on spiral galaxy and determine whether it is seen end-on or side-on, the introduction of dust in the simulations has allowed us to develop criteria to determine in which half of the galaxy the near side of the bar is located. Of course, the distribution of dust in real galaxies will inevitably be more complex than the highly idealised distribution adopted here. Nevertheless, the features due to dust in the PVDs of Figure~\ref{fig:dust} can probably still be used as a guide to interpret asymmetries present in real data. \section{Discussion\label{sec:discussion}} \nopagebreak \subsection{Bar Diagnostics\label{sec:diag}} \nopagebreak Unlike periodic orbits studies (\markcite{km95}KM95; \markcite{ba99}Paper~I), where one has to adopt a way of populating the orbits, hydrodynamical simulations provide both the velocity and density of the gas. In particular, if some periodic orbit families intersect, it is not necessary to make a choice between them, because the simulations will reveal which family the gas follows in each region (and to what extent). Nevertheless, as we will discuss later in this section, the comparison with observations can still present some problems, since observations involve the strength of a given emission line rather than the gas density. Notwithstanding the problem of populating the orbits, there is generally a good agreement between the PVDs obtained from periodic orbit calculations in \markcite{ba99}Paper~I and those obtained here from hydrodynamical simulations. This is because there are several similarities between the periodic orbits structure of the models and the gas flow (\markcite{a92b}A92b). We have in both cases a central PVD component, which we identified in \markcite{ba99}Paper~I with the $x_2$ orbits and in this paper with the nuclear spiral. Further out, we have the domain of the $x_1$ orbits, which covers the outer (or entire) bar regions of the simulations. In \markcite{ba99}Paper~I, by studying in detail the signatures of various periodic orbit families in the PVDs, we gained useful insight into the projected structure and kinematics of the gas from first principles. This allowed us to obtain a deeper understanding of the PVDs produced in the hydrodynamical simulations. The main feature of the PVDs is the gap (fainter region), present at all viewing angles, between the signature of the nuclear spiral and that of the outer parts of the simulations. Such a structure would not be possible in an axisymmetric galaxy and it unmistakably reveals the presence of a bar or oval in an edge-on disk. This gap occurs because of the large-scale shocks which are present in bars, and which drive an inflow of gas toward the center, depleting the outer bar regions. The gaps present in the PVDs produced with periodic orbits (see \markcite{km95}KM95; \markcite{ba99}Paper~I) are different in nature from the ones observed here, being mainly due to the absence of populated orbits in certain regions of the models, particularly near corotation. As a result, in these studies, the low density region extends well beyond corotation, although its exact extent depends on which orbits are neglected (e.g.\ as self-intersecting) and how the other orbits are populated. For example, in Figure~1 of \markcite{km95}KM95, the gap extends to almost twice the corotation radius. We recall that 3D $N$-body simulations of rotating disks produce bars which, when viewed edge-on, appear boxy-shaped if seen end-on and peanut-shaped if seen side-on (see, e.g., Combes \& Sanders \markcite{cs81}1981; Combes et al.\ \markcite{cdfp90}1990; Raha et al.\ \markcite{rsjk91}1991). Taking the maximum height of the peanut-shaped bars to occur at half the corotation radius (Combes et al.\ \markcite{cdfp90}1990), we find that, in the \markcite{km95}KM95 case, the ratio of the radial extent of the gap in the PVDs to the radius where the maximum height of the peanut-shaped bulges occurs should be approximately 4 (or slightly less because of projection effects, associating the bulges with edge-on bars). On the other hand, in the hydrodynamical simulations presented here, the low density region of the PVDs reflects the low density region in the outer parts of the bar, and should thus be within corotation. When a bar is seen side-on, the ratio of the extent of the gap to the radius of maximum height of the peanut-shaped bulge should thus be approximately 2 (or slightly less). \markcite{km95}KM95 report that this ratio is about 2 for the two galaxies they studied, in agreement with the prediction of our hydrodynamical simulations. The same can be said for the galaxies studied by Bureau \& Freeman (\markcite{bf97}1997) and \markcite{bf99}BF99. These results can only be reconciled with the periodic orbits approach if the maximum height of the peanut-shaped bulges occurs near corotation (rather than halfway). As we pointed out in \S~\ref{sec:001}, the shapes of the features present in the PVDs vary with the viewing angle, and can thus in principle be used to constrain that quantity in an observed galaxy. The envelope of the signature of the outer bar region in the PVDs reaches very high radial velocities (compared to those in the outer parts of the simulations) for viewing angles close to the bar, and relatively low velocities for viewing angles perpendicular to it. However, the envelope is so faint in most cases that it is unlikely to be of much use with real data (see \S~\ref{sec:others}). The signature of the nuclear spiral has an inverted S-shape for some viewing angles and is almost solid-body for others. That feature is rather thick and does not show much fine structure, however, so it may be hard to use in conjunction with observations. The best viewing angle diagnostic is probably provided by the ratio of the maximum radial velocity reached by the nuclear spiral component to the velocity in the outer parts of a galaxy. This ratio should be greater than unity for viewing angles roughly perpendicular to the bar and smaller than unity for viewing angles parallel to it. Analysing their sample, \markcite{bf99}BF99 find that the above is in good agreement with their results, provided they make the reasonable assumption that bars are peanut-shaped when seen close to side-on and boxy-shaped when seen close to end-on (see, e.g., Combes \& Sanders \markcite{cs81}1981; Combes et al.\ \markcite{cdfp90}1990). Other properties of the PVDs, such as the maximum radial velocity reached or the faintness of the signature of the outer (or entire) bar regions, can also help to constrain the mass distribution and bar properties of an observed galaxy (see \S~\ref{sec:others}). In practice, however, the analysis and modeling of spectroscopic data should not be a trivial task. If a barred system has no nuclear spiral (or, equivalently, does not have an extended $x_2$ family of periodic orbits), there will not be a nuclear spiral signature in the PVDs and no gap, or double-peaked structure. In addition, the surface density in the bar region can be extremely low. An example of such a model was discussed in \S~\ref{sec:086} (model 086). In such cases, the only component likely to be detected observationally is the solid-body signature of the outer parts of the galaxy, and the kinematical detection of the bar will not be straightforward. The first step would be to rule out such a slowly rising rotation curve (the rotation curve being defined as the upper limit of the envelope of an observed PVD), for example by calculating the shape of the rotation curve expected from surface photometry. Unfortunately, this would only show that the type of gas observed (ionised, neutral, or molecular) is not present in large quantities in the central region of the galaxy, but it would not provide any information on the cause of this depletion. It is thus probably necessary to use stellar kinematics to identify a bar in such systems. Because a large percentage of the stars in the central regions of barred disks are expected to be trapped around the $x_1$ orbits, there should be a clear bar signature in the PVDs. We shall explore possible diagnostics based on the stellar kinematics (and corresponding PVDs) in \markcite{ab99}Paper~III. Cases with no nuclear spiral component should be a minority, however, at least among strongly barred early-type spiral galaxies, since observational evidence argue that these systems possess ILRs (see, e.g., Athanassoula \markcite{a91}1991; \markcite{a92b}A92b). In particular, the fact that the dust lanes in most early-type strongly barred galaxies are offset along the leading sides of the bar, rather than centered and close to its major axis, is a strong argument for the existence of ILRs (\markcite{a92b}A92b). In addition, out of 17 galaxies with a boxy/peanut-shaped bulge in the sample of \markcite{bf99}BF99, only four have no nuclear component in their PVD. The lack of a nuclear component in these four galaxies could be due either to the lack of ILRs, or to the lack of emitting gas around the ILRs region. At least 13 out of 17 galaxies, i.e. an overwhelming majority, have ILRs. Nevertheless, the existence of ILRs deserves further study, particularly for later type spirals. This can be done, for example, by high resolution kinematical studies, to show the change of direction of the orbits or streamline ellipticities in the central parts of galaxies (e.g.\ Teuben et al.\ \markcite{tsaa86}1986), or by calculating the families of periodic orbits in galaxy potentials derived from observations, to show the existence or non-existence of the $x_2$ (and $x_3$) family. \subsection{NGC~5746-Like PVDs\label{sec:ngc5746}} \nopagebreak Although our simulations cover a fair fraction of the available parameter space (see Table~1 in \markcite{a92a}A92a), we have never come across a PVD showing a clear ``figure-of-eight'', as suggested by \markcite{km95}KM95. In other words, the upper envelope of the low density region of the simulated PVDs is never as pronounced as it is in NGC~5746 or, to a lesser extent, in NGC~5965 and NGC~6722 (see \markcite{km95}KM95; Bureau \& Freeman \markcite{bf97}1997; \markcite{bf99}BF99). Based on periodic orbit calculations, we expect this region of the PVDs to originate from material near the ends of the bar on its {\em minor} axis (see, e.g., Fig.~2b in \markcite{ba99}Paper~I). Some models, particularly $n=0$ models, indeed have secondary density enhancements near the edge of the bar, just outside the offset shocks which we have already discussed (see Fig.~3 in \markcite{a92b}A92b). These enhancements are due to the small distance between the outer ILR and the tip of the $x_1$ family characteristic curve in the characteristic diagram (\markcite{a92b}A92b). This leads to orbit crowding, particularly near the bar minor axis. In these models, the upper envelope of the gap in the PVDs is stronger than that in most other models. \placefigure{fig:088} Figure~\ref{fig:088} shows two density distributions for model 088 and the corresponding PVDs for a viewing angle $\psi=45\arcdeg$. On the left (Fig.~\ref{fig:088}a,b), the entire density distribution and PVD are displayed. On the right (Fig.~\ref{fig:088}c,d), most of the {\em PVD} was masked out, leaving only the region of interest in objects like NGC~5746. The corresponding density distribution was obtained using a simple inversion scheme. The strong upper envelope of the low density region of the PVD is clearly due to the secondary density enhancement on the leading side of the bar. A similar effect is observed in many $n=0$ models with low bar axial ratio $a/b$, small Lagrangian radius $r_L$, and/or low central concentration $\rho_c$ (e.g.\ models 013-016, 040-043, 061-063, 087-088, and 108-110). Although the intensities obtained in these simulations are much lower than the strong envelopes observed in objects like NGC~5746, we must remember that it is the density which is plotted here, and the gaseous emission in the secondary density enhancements could be somehow enhanced, e.g.\ through shocks. Features similar to those density enhancements are observed in real barred galaxies, where ``plumes'' are sometime seen on the leading sides of bars (in NGC~1365 for example). We suggest here that these plumes may be at the origin of the strong upper envelope of the gap observed in the PVD of some galaxies. We must stress, however, that a strong envelope is unusual; it is absent from the PVD of most galaxies. In the sample of \markcite{bf99}BF99, only 2 galaxies out of 17 present clear indications of such a feature. This seems in agreement with the explanation presented above, as strong ``plumes'' are not often observed. \subsection{Limitations of the Models\label{sec:limits}} \nopagebreak When using the bar diagnostics developed in this paper to interpret observational data, one has to take into account the fact that the PVDs were calculated using the gas {\em density} in the simulations and not the strength of a given gaseous emission line, which is what one usually observes. Physical conditions in the gas vary across the simulated disks, and will lead to different excitation mechanisms dominating in different parts of the disks. Regions of high density and low shear are likely to have higher star formation rates and thus more photoionised gas than elsewhere. Conversely, the gas in or near shocks, such as the nuclear spiral, will be mainly shock-excited (see Binette, Dopita, \& Tuohy \markcite{bdt85}1985; Dopita \& Sutherland \markcite{ds96}1996), with possible star formation depending on the shear. Because different excitation mechanisms lead to different emission line ratios, the relative amplitude of the various components of a PVD (e.g.\ the intensity of the signature of the outer parts of the disk versus that of the nuclear spiral) will depend on the emission line used in the observations. For example, Bureau \& Freeman (\markcite{bf97}1997) and \markcite{bf99}BF99 found that, for many objects, the signature of the nuclear spiral was very strong in the [\ion{N}{2}]~$\lambda$6548,6584 lines but was almost absent in H$\alpha$, probably indicating that it is shock-excited (the other components of the PVDs showed [\ion{N}{2}]~$\lambda$6584/H$\alpha$ ratios typical of \ion{H}{2} regions). The PVDs were not corrected for stellar absorption, however, so these results should be interpreted with caution. Nevertheless, the PVDs produced should only be used as guide when interpreting kinematical data, and one should have a basic understanding of the mechanisms involved in the production of a given line before using the PVDs for comparison purposes. We stress that the morphology of the PVDs (the multiple components) is a more significant bar diagnostic than the distribution of intensity (the relative amplitude of the components). Conversely, under certain assumptions about the physical conditions in the gas, it is possible to use the observed emission line ratios (e.g.\ H$\alpha$/H$\beta$) to measure the extinction due to dust in data. This could prove useful to interpret asymmetries present in observed PVDs (see \S~\ref{sec:dust} for a more complete discussion of likely dust effects). Contrary to the ``building blocks'' approach of \markcite{ba99}Paper~I, which used combinations of periodic orbit families to model the structure and kinematics of barred galaxies, the hydrodynamical simulations presented here inherently take into account the collisional nature of gas, so the kinematics of the gaseous component is more accurately modeled. Approximations to the gas properties are nevertheless necessary, and we recall that the gas was treated as ideal, isothermal, infinitely thin, and non-self-gravitating. How much do our results depend on these assumptions, or, in other words, how model dependent are they? The interstellar medium is a complicated multi-phase mixture, which can only be described schematically, particularly in non-local studies covering an object the size of a galaxy. Two different approaches have been developed so far. In the first one, the gas is treated as ballistic particles which, when they collide, lose energy according to pre-specified recipes (e.g.\ Miller, Prendergast, \& Quirk \markcite{mpq70}1970; Schwarz \markcite{s81}1981, \markcite{s84}1984; Combes \& Gerin \markcite{cg85}1985). Unfortunately, the results of these simulations can depend on the adopted collision law (Guivarch \& Athanassoula \markcite{ga99}1999), at least as far as the shocks in the barred region are concerned. In the second approach, a hydrodynamical treatment is used, solving the Euler equations with the help of a grid (e.g.\ van Albada \& Roberts \markcite{vr81}1981; van Albada, van Leer, \& Roberts \markcite{vvr82}1982; Mulder \markcite{86}1986; Piner, Stone, \& Teuben \markcite{pst95}1995), Smooth Particles Hydrodynamics (SPH; Lucy \markcite{l77}1977; Gingold \& Monaghan \markcite{gm77}1977; Hernquist \& Katz \markcite{hk89}1989; etc), or beams (Sanders \& Prendergast \markcite{sp74}1974). A number of these studies assume an isothermal equation of state, following the model of Cowie \markcite{c80}(1980), who calculated the equation of state for an ensemble of clouds and found it could be described as isothermal, provided the clouds have an equilibrium mass spectrum. Comparing the results of all these schemes is beyond the scope of this paper. In all approaches, the more reliable codes produce shocks in the bar region which are more or less offset from the bar major axis towards its leading sides. These shocks result in an inflow of gas and a substantial lowering of the density in the outer bar region. Although the precise location, shape, and persistence of the shocks differ, those are relatively small effects, compared to the fact that we weight by density rather than by the strength of an emission line. The main reason for adopting the present hydrodynamical code is that it gave very good results in many previous studies (see \markcite{a92b}A92b). Because dissipation ensures that the gas layer in spiral galaxies remains thin, the two-dimensional nature of the simulations should not be a factor limiting the applicability of the results to the interpretation of real data. Furthermore, vertical motions have no direct consequence on the PVDs produced as the movement is perpendicular to the line-of-sight. It is somewhat harder to gauge the consequences of the fact that we have ignored the gas self-gravity. Because our mass model represents best (barred) early-type spirals (see \markcite{a92a}A92a), where the fraction of the total mass in gas is typically less than 10\%, the contribution of the gas to the potential and large scale forces is negligible. Indeed, Lindblad, Lindblad, \& Athanassoula (\markcite{lla96}1996) found that including self-gravity in their model of NGC~1365 made very little difference to the global picture. It may well be, however, that near high density structures such as the nuclear spiral, the self-gravity of the gas is important. \section{Summary and Conclusions\label{sec:conclusions}} \nopagebreak Our main aim in this paper, the second in a series, was to develop diagnostics to identify bars in edge-on spiral galaxies using the particular kinematics of the gaseous component of barred disks. To achieve this goal, we ran two-dimensional hydrodynamical simulations of the gas flow in the potential of a barred spiral galaxy mass model. We constructed position-velocity diagrams (PVDs) from those simulations, using an edge-on projection and various viewing angles with respect to the major axis of the bar. The presence of shocks and inflows in the simulations allowed us to develop better bar diagnostics than those presented in our first paper (Bureau \& Athanassoula \markcite{ba99}1999), based on periodic orbits calculations. We analysed in detail two simulations which are prototypes of simulations for models with and without inner Lindblad resonances (which we associate with the existence of $x_2$ periodic orbits). We showed that, for models allowing $x_2$ orbits, the nuclear spiral which is created in the center of the simulations produces a strong inverted S-shaped signature in the PVDs. This signature reaches high radial velocities (compared to those in the outer parts of the simulations) when the bar is seen side-on, and relatively low velocities when the bar is seen end-on. The flow in the outer bar region (the entire bar region if no nuclear spiral is present) produces a parallelogram-shaped signature in the PVDs, being associated with the $x_1$ periodic orbits. Because the flow is mostly along the bar in that region, the highest velocities are now reached for viewing angles close to the bar major axis. In the outer parts of the simulations, the flow is almost circular and produces a strong almost solid-body signature in the PVDs for all viewing angles. Shocks within the bar are present in most simulations and lead to a depletion of the gas in the region of the bar occupied by an $x_1$-like flow. Thus, if a nuclear spiral is present, a bar can easily be identified in an edge-on spiral galaxy, as there will be a gap in the PVD between the signature of the nuclear spiral and that of the outer parts of the galaxy. If there is no nuclear spiral, it may still be possible to detect a bar, but only with the help of photometry and/or stellar kinematics. The envelope of the signature of the nuclear spiral, and to a lesser extent that of the outer bar region, is most useful to determine the orientation of a bar with respect to the line-of-sight. It is nevertheless hard to discriminate between two viewing angles on either side of the bar. We showed that adding dust to the simulations helps break this degeneracy. We also produced PVDs for a range of simulations covering most of the fraction of parameter space likely to be occupied by real galaxies. These simulations can be used to constrain the mass distribution and bar properties of an observed system. In particular, the presence or absence of the signature of a nuclear spiral in a PVD places strong constraints on the values the parameters of our mass model may take. The nuclear spiral can be absent for high bar axial ratios and/or bar quadrupole moments, and for low Lagrangian radii and/or central concentrations. \acknowledgments We thank K.\ C.\ Freeman, A.\ Bosma, and A.\ Kalnajs for comments on the manuscript and J.-C.\ Lambert for his computer assistance. E.\ A.\ thanks G.\ D.\ Van Albada for making available to her his version of the FS2 code. M.\ B.\ acknowledges the support of an Australian DEETYA Overseas Postgraduate Research Scholarship and a Canadian NSERC Postgraduate Scholarship during the conduct of this research. M.\ B.\ would also like to thank the Observatoire de Marseille for its hospitality and support during a large part of this project.
\section{Introduction} The generation of pure states of the electromagnetic field is an important issue in quantum optics. Nonclassical states such as squeezed \cite{slus} and Schr\"odinger cat states \cite{brun} have been already generated, and several schemes for the generation of arbitrary quantum states, named quantum state engineering, have been proposed throughout. Normally those methods have the vacuum state $|0\rangle$ as a starting point for the field. The energy necessary to build up a given state may be supplied by atoms \cite{schl}, by coherent plus one-photon fields \cite{wels}, as well as classical pumps \cite{eber,vidi}. In any case, photons are coherently added to an \underline{already pure state} (vacuum). It would be therefore interesting to verify how quantum state generation could be achieved if we depart from less favourable initial conditions, i.e., with initial mixed states, instead. This problem has already been addressed in Ref. \cite{moya}, where it is described a scheme of purification of thermal fields by means of one photon-transitions. The central point of the method is the progressive transfer of coherence from atoms to the cavity field. Atoms are prepared in superpositions of circular Rydberg states (while crossing a Ramsey zone apparatus), and successively injected into a high-Q cavity, where the field, resonant with the atomic transition, is built up. The transfer of coherence from conveniently prepared atoms to cavity fields consists in an important mechanism for the investigation of quantum aspects of light and matter. It also leads to interesting effects, such as the atomic population trapping \cite{zahe,jona}, for instance. Besides, it is comparatively easier to prepare coherent superpositions of atomic states. In this paper we are going to be concerned with the purification of mixed states in a two-photon micromaser. In our scheme, conveniently prepared three-level atoms undergoing two-photon transitions are injected into a cavity. Obviously, we expect such a scheme to lead only to certain types of pure states, because of the effective two photon interaction. Nevertheless, because for the same reason, as we are going to show, the cavity field purification may be attained faster than in a one-photon scheme. Moreover, in a two-photon micromaser, a considerably higher degree of purification is achieved. This paper is organized as follows; in section 2 we present our scheme of field generation in section 3 we discuss the statistical properties of the produced fields; in section 4 we show the evolution of the field according to the phase space representations; in section 5 we summarize our conclusions. \section{General scheme} We consider three-level atoms in a ladder-type configuration. The intermediate level may be adiabatically eliminated, resulting the following effective Hamiltonian \cite{simon} \begin{equation} \hat H=\hbar \omega \hat a^+ \hat a + \frac{1} {2} \hbar (\omega_0+\chi \hat a^+ \hat a) \sigma_3 + \hbar \lambda (\hat a^{+2} \sigma_- + \hat a^2 \sigma_+), \end{equation} where $\lambda$ is the atom-field coupling constant; $\omega$ is the field frequency; $\omega_0$ the atomic transition frequency; $\sigma_3, \sigma_- ,\sigma_+$ are the atomic operators; $\hat a^+$, $\hat a$ the field operators, and the detuning $\Delta=\omega_0- 2\omega$. The Stark shift coefficient is $\chi$. The evolution operator relative to the Hamiltonian above is straightforwardly obtained \cite{sten}, resulting \begin{equation} \hat{U}(t)= \left[ \begin{array}{cc}\hat{\alpha}_n(\gamma)& \hat{\beta}_n(\gamma)\\ \hat{\beta}_n(\epsilon) & \hat{\alpha}_n^{\dagger}(\epsilon) \end{array} \right], \end{equation} where \begin{equation} \hat{\alpha}_n(\gamma) = \cos(\hat \gamma_n \lambda t)+i \frac{\sin(\hat \gamma_n \lambda t)} {\hat \gamma_n} \left(\frac{\frac{\Delta}{2} + \chi \hat n} {\lambda}\right) \end{equation} \begin{equation} \hat{\beta}_n(\epsilon) = i \frac{\sin (\hat \epsilon_n \lambda t)} {\hat \epsilon_n} \end{equation} with \begin{equation} \hat \gamma_n^2=\left(\frac{\frac{\Delta}{2} + \chi \hat n}{\lambda}\right)^2 +(\hat n+1)(\hat n+2), \end{equation} and \begin{equation} \hat \epsilon_n^2=\left(\frac{\frac{\Delta}{2} + \chi \hat n}{\lambda}\right)^2 +\hat n (\hat n-1). \end{equation} The time evolution of the total (atom-field) density operator $\hat{\rho}^{af}$ will be \begin{equation} \hat{\rho}^{af}(t)=\hat{U}(t)\hat{\rho}^{af}(0)\hat{U}^\dagger(t). \end{equation} The successively injected atoms are prepared, before entering the cavity, in the following superposition of upper and lower levels \begin{equation} |\psi\rangle=a|g\rangle+be^{i\phi}|e\rangle, \end{equation} being $a$ and $b$ real (nonzero) numbers and $\phi$ a relative phase. Therefore the atom will be, at $t=0$, in the pure state $\hat{\rho}^{a}(0)=|\psi\rangle \langle\psi |$, while the field is in a mixed state $\hat{\rho}^{f}(0)$. We assume that the total initial state may be factorized, or $\hat{\rho}^{af}(0)=\hat{\rho}^{a}(0)\otimes\hat{\rho}^{f}(0)$. As usual, the field state is readily obtained by tracing over the atomic variables, or $\hat{\rho}^{f}(t)=Tr_a\left[\hat{\rho}^{af}(t)\right]$. After injecting $N$ atoms, the matrix elements of the field density operator in the number state basis $\langle n|\hat{\rho}^{f}|n'\rangle=\rho_N ^f(n,n')$ will obey the following (micromaser) recurrence formula \begin{eqnarray} \rho_N ^f(n,n')&=&[b^2 \alpha_n(\gamma)\alpha_{n'}^{\dagger}(\gamma) +a^2\alpha_n(\epsilon) \alpha_{n'}^{\dagger}(\epsilon)]\rho_{N-1} ^f(n,n')\nonumber\\ &+&a^2\beta_n(\gamma)\beta_{n'}(\gamma)\sqrt{(n+2)(n+1)(n'+2)(n'+1)} \rho_{N-1} ^f(n+2,n'+2)\nonumber\\ &+&b^2 \beta_n(\epsilon)\beta_{n'}(\epsilon)\sqrt{n(n-1)n'(n'-1)} \rho_{N-1} ^f(n-2,n'-2)\nonumber\\ &+&iab e^{i\phi}\alpha_n(\gamma)\beta_{n'}(\gamma)\sqrt{(n'+2)(n'+1)} \rho_{N-1} ^f(n,n'+2)\nonumber\\ &+&iab e^{-i\phi}\alpha_n(\epsilon)\beta_n(\epsilon)\sqrt{n'(n'+1)} \rho_{N-1} ^f(n,n'-2)\nonumber\\ &-&iab e^{-i\phi}\beta_n(\gamma)\alpha_{n'}^{\dagger}(\gamma)\sqrt{(n+2)(n+1)} \rho_{N-1} ^f(n+2,n')\nonumber\\ &-&iab e^{i\phi}\beta_n(\epsilon)\alpha_{n'}^{\dagger}(\epsilon) \sqrt{n(n-1)}\rho_{N-1} ^f(n-2,n').\label{heq} \end{eqnarray} From this matrix elements which represent the state of the field, we are able to determine under which conditions we may attain a reasonable purification of the field, namely, departing from a mixed state. We note that because $a$ and $b$ are both nonzero, off-diagonal elements (number state basis) will be generated, characterizing the transfer of coherence. The time-dependent field purity parameter $\zeta$, defined as \begin{equation} \zeta=1-Tr\left[(\rho^f){}^2\right]=1-\sum_{n,n'}|\rho^f_N(n,n')|^2, \end{equation} will provide us the necessary guidance in order to purify the initial field. If $\zeta=0$, this means that $\hat{\rho}$ represents a pure state. Therefore, we now seek optimum conditions for field purification, which corresponds to values of $\zeta$ as close to zero as possible. It would be also interesting to build up the field, i.e., to increase the mean energy of the field while the atoms cross the cavity. Here we are going to consider two different initial fields, the thermal state (mean photon number $\overline{n}$) \begin{equation} \hat \rho^f_{th}(0)=\sum_{n=0}^\infty\frac{\overline{n}^{n}}{(\overline{n}+1)^{n+1}} |n\rangle \langle n|,\label{ther} \end{equation} and the mixed (phase diffused) coherent state \begin{equation} \hat \rho^f_{co}(0)=\sum_{n=0}^\infty \frac{|\alpha|^{2n} e^{-|\alpha|^2}}{n!} |n\rangle \langle n|.\label{coher} \end{equation} The overall features of the process are similar in either case, but important differences concerning the statistical properties of the field arise during the evolution. \section{Results} First we inspect the time-evolution of the field purity for one atom, choosing an interaction time that leaves the field purer as the atom exits the cavity. Thus the next atom entering the cavity will interact with a `less mixed field'. In figure 1 we have plots of $\zeta$ as a function of time, for an initial thermal field with $\overline{n}=10$, after having passed 1, 20 and 100 atoms. We note that for several ranges of times, the field is purer than the initial (mixed) state. In order to choose an optimum interaction time, we examined the steady state, when $N$ atoms had already crossed the cavity, and imposing the condition that the final field state should be considerably purer than the initial state. A second constraint is that the mean photon number inside the cavity should either increase or remain constant. For simplicity we have chosen the same interaction time for all atoms. This was the main guidance for choosing that time. After calculating the field evolution for a range of times, we found that the optimum interaction time turns out to be $T\approx 12.2/\lambda$. The atoms are assumed to be prepared in a equally weighed state ($a=b=1/\sqrt{2}$), $\Delta=\chi=\lambda$, and $\phi=0$. In figure 2 we have a plot of the field purity ($\zeta$) of the final state, as a function of the (scaled) interaction time $\lambda T$. We note that the field purity is maximum ($\zeta\approx 0$) for certain times. However, this corresponds to the case in which photons are subtracted from the cavity field, leading to either the vacuum state or the one-photon state. This is seen in figure 3, where we have the mean photon number of the cavity field as a function of the interaction times. The dashed line indicates the interaction time $T\approx 12.2/\lambda$ we have chosen for our scheme. After fixing the interaction time, we may now analyze how the field changes as the atoms successively enter the cavity. In figure 4 we have a plot of $\zeta$ as a function of the number of atoms, for both thermal and mixed coherent fields. We note that in either case the increase of purity (decrease of $\zeta$) is substantial even for $N=100$ atoms, and saturation ($\zeta\approx 0.53$) already exists around $N=200$ atoms, i.e., further injection does not improve the situation. We note that a larger degree of purity is achieved faster for a phase diffused coherent state than for the thermal state, as we see in figure 4. This may be understood if we compare both initial photon number distributions with the distribution of the steady state field. Although there are not non-diagonal elements in the density matrix (number state basis) for any of the initial fields, the phase-diffused coherent state has a more symmetrical (Poissonian distribution) than the thermal state (geometrical distribution). In figure 5 we have a plot of the photon number distribution of the steady state for an initial thermal field. We note that it has the overall shape similar to a Poissonian, the distribution of the phase diffused coherent state, apart from the strong oscillations. It is therefore reasonable to expect the final state to be achieved more easily in that case, rather than for a thermal state, result which is apparent in the purity curve. In figure 6 it is shown the mean photon number $\langle\hat{n}\rangle\equiv \sum \rho^f_N(n,n)n$ in the cavity as a function of the number of atoms. We note the increase of energy from the initial $\langle n \rangle=10$ up to $\langle n \rangle\approx 32$ photons, and occurs in a similar way for both thermal and mixed coherent initial fields. Another interesting aspect is that the generated fields not only become purer than the initial ones, but are also nonclassical in the sense that they may display sub-Poissonian statistics and/or anti-bunching. This fact is represented in figure 7, where Mandel's Q parameter, defined as $Q=\Delta\hat{n}^2/\langle n \rangle-1$ is plotted as a function of the number of atoms $N$. We note that for both an initial thermal field (figure 7 (a)), and for a mixed coherent state ((figure 7 (b)), the field becomes sub-Poissonian after around 100 atoms have crossed the cavity. However, there is an important difference between the two cases for a not so large number of atoms. For an initial thermal state the field becomes monotically less super-Poissonian, while for an initial mixed coherent state (it is Poissonian at $t=0$) it starts becoming super-Poissonian, i.e., the parameter Q grows until it reaches a maximum value (after passing around 30 atoms). Then it starts decreasing, turning sub-Poissonian even faster than the thermal state case, as it is shown in figure 7. In both cases the field becomes anti-bunched after injecting more than 100 atoms. The generated field also displays strong oscillations in its photon number distribution (see figure 5) which characterizes Schr\"odinger cat-like states \cite{vidi1}. We could then seek for more details about the generated state. Because we have obtained the density operator in the number state basis, it is convenient to switch to other representations, such as the phase-space quasiprobabilities. \section{Phase-space approach} Quasiprobability distributions have become important tools not only for quantum state characterization, but have also been playing an active role in quantum state reconstruction \cite{ulf}. Here we are going to be concerned with the characterization of the field produced in the cavity. Specially useful in this case is the series representation of quasiprobabilities \cite{moya1} \begin{equation} P(\beta;s)=\frac{2}{\pi}\sum_{k=0}^\infty(-1)^k\frac{(1+s)^k}{(1-s)^{k+1}} \langle\beta,k|\hat{\rho}^f_N|\beta,k\rangle. \end{equation} We may represent the field density operator as $\hat{\rho}^f_N=\sum_{n,n'} \rho^f_N(n,n')|n\rangle\langle n'|$, where the matrix elements $\rho^f_N(n,n')$ are the ones in equation \ref{heq}. For $s=1$ we have the Q function, and $s=0$ we obtain the Wigner function. In figure 8 we have a plot of the contours of the Q function after passing $N=200$ atoms through the cavity, for a field initially prepared in a thermal state. We note that four peaks are formed around the origin, indicating that a certain symmetry of the quasiprobabilities around the origin is preserved during the generation process. We would also like to mention that with a convenient choice of parameters, in such a way that photons are subtracted, we may also reach a pure field, e.g., a one-photon state. Nevertheless this case is not as interesting as if we are able to purify the field and build it up at the same time, as we have shown above. The quasiprobabilities approach strongly suggests that the steady-state field is constituted by a superposition of four deformed Gaussian-like distributions, which resemble squeezed state's ones. We may therefore conjecture that the resulting state is in fact some kind of superposition of four squeezed coherent states. This is also supported by the fact that the photon number distribution displays strong oscillations, characteristic of those kind of superpostions \cite{vidi2}. Because our method is based on two-photon interactions, which are necessary for squeezed state generation, we could have expected the generation of fields somehow related to squeezed states. We would like to remark that our aim was to find a way of allowing the generation of the purest possible state, even departing from highly mixed states, instead of establishing a more general quantum state engineering scheme. \section{Conclusions} We have investigated the process of generation of a nonclassical state departing from a thermal state by means of a two-photon micromaser. A reasonable degree of field purity is quickly achieved, and the generated field presents nonclassical properties such as sub-Poissonian character. Therefore the purification procedure also leads to the generation of a nonclassical state. We should remark that after passing about only $N=100$ atoms in the cavity, we were able to obtain a degree of purity ($\zeta\approx 0.53$) higher than in one-photon transitions ($\zeta_{op}\approx 0.65$) \cite{moya}. Of course the more intense the initial field is, the more difficult will be to purify it as time goes on. Here we have attained a good degree of field purity even having started with a relatively noisy field, containing around $\overline{n}=10$ thermal photons. We have studied the case in which the mean number of photons of the field is increased. We have so far identified two types of states; the four-peaked distribution shown in figure 8, and the trivial case (photons are subtracted), which leads to either to the vacuum state or the one-photon state. In both cases the states have a representation in phase space symmetric in relation to the origin, similarly to the initial states. We have neglected losses as an approximation, and of course we expect decay to compete against the generation process. Nevertheless we would like to make a couple of remarks; the total estimated time for the experiment is around five times longer than the energy decay time in a state-of-the-art high $Q$ cavity. In the beginning of the process, one-photon losses are not expected to be of much importance, given the initial states, which will not have their statistics substantially changed by decay. However, as times goes on, decay will surely not favour the process. Nevertheless the investigation of ideal situations surely enlightens the discussions on the question of quantum state generation. \newpage \noindent {\bf\LARGE Acknowledgements} \vspace{0.3cm} Two of us, AFG and JAR, would like to acknowledge financial support from Conselho Nacional de Desenvolvimento Cient\'\i fico e Tecnol\'ogico (CNPq), Brazil. \newpage
\section{Introduction} Phenomena strikingly deviating from the conventional understanding can take place if the speed of photons is allowed to vary from the constant light velocity $c$. Coleman and Glashow (1997) argued that the velocity $c_0$ of light in the vacuum in a frame moving relative to the rest frame of the universe can differ from the maximal attainable speed $c$ of a material body. If $c_0 > c$, the photon 4-momentum $(\vec E /c_0, E)$ is time-like, allowing a sufficiently energetic photon to decay into an electron-positron pair. Alternatively, if $c_0 <c$, a charged particle will lose energy via vacuum \v Cerenkov radiation because the particle speed approaching $c$ at high energies can exceed the light speed $c_0$. Thus, the distance that energetic $\gamma$-rays or charged particles can travel through is shortened for either $c_0 > c$ or $c_0 < c$. However, the violation of relativity may also have a possibility of increasing the travelling distance of high energy radiation rather than decreasing. The speed of photons can vary due to modifications of the vacuum (Latorre et al. 1995). Different energies excite the vacuum differently, and with quantum gravity, Amelino-Camelia et al. (1997) speculated that the energy-momentum relation for photons is modified to violate Lorentz invariance and the anomalous effect increases as photon energy approaches the Planck energy scale ($ \sim 10^{19}$ GeV). Amelino-Camelia et al. (1998) has proposed a test for the modified speed by measuring the arrival time structure of photons from objects at cosmological distances. The propagation speed is $dE / dk = c (1 - \xi _{\gamma} E / E_0)$ from the modified relation of energy $E$ and momentum $k$, \begin{equation} c^2 k^2 = (E + \xi _{\gamma} {E^2 \over 2 E_0})^2 = E^2 + \xi _{\gamma} {E^3 \over E_0} \end{equation} with $\xi _{\gamma} = \pm 1$ and $E_o \sim 10^{19}$ GeV, according to Amelino-Camelia et al. 1998. By assuming the existence of a preferred frame for Lorenz transformation but preserving other kinds of invariance, Coleman and Glashow 1998 gives a relation, which is similar to (1) and equivalent to $c^2 k^2 = E^2 - m_i^2 c^4+ (c^2-c_i^2) E^2$ for $ck \approx E \gg m_i c^2$, where $c_i$ is the maximal attainable speed of particle species $i$ with mass $m_i$. Our study in what follows will be in an explicit form based on (1), but the results are, to some extent, applicable also to the other case by putting $c^2-c_i^2$ in the place where the term $\xi _i E/E_0$ appears. The electromagnetic radiation of the shortest wavelength so far detected is TeV $\gamma$-rays (see for example, Weekes et al. 1997; Ong 1998), but even for these high energy photons, the effect of velocity variation remains as small as a fraction $\sim 10^{-15}$ of the light velocity. However, a higher degree of anomalous effect is expected for the absorption process, in which $\gamma$-rays are converted into an electron--positron pair when they collide with soft photons of the universal radiation field. Photons of TeV energies react with infrared photons, and PeV $\gamma$-rays with the 2.7K CMB (cosmic microwave background). From the number density of the photons of the universal infrared and CMB, the collision mean free path is usually estimated as about 100 Mpc and 10 kpc ({\it e.g.} Strong et al. 1974; Stecker et al. 1992), respectively for TeV and PeV $\gamma$-rays. The largest distance we can see with the high energy radiations is thus much shorter than those of X-rays and optical photons of longer wavelengths. \section{Kinematical constraints on the photon-photon collision process} The threshold energy is determined by a function of the total energy $W$ and momentum $Q$ of the initial state, $W^2 - c^2Q^2$, which must exceed $(2 mc^2)^2$ for the pair creation, where $m$ is the electron mass. In the initial state of a $\gamma$-ray obeying the relation (1) and soft photons of energy $\varepsilon$, the square of the energy $E$ and the momentum $\sim E/c$ in $W^2 - c^2Q^2$ almost cancel each other to leave $4\varepsilon E - \xi _{\gamma} E^3/E_0$ (the two photons are assumed to be from opposite directions). The remaining two terms are, when $\varepsilon / E$ is as small as $\xi_{\gamma} E / E_0$, of similar magnitude to each other, incurring a considerable degree of anomalous effect. In addition, the two terms become as large as $m^2 c^4$ when $E \approx 10^{12} \xi_{\gamma} ^{-1/3} \approx 10^{12} \varepsilon ^{-1}$ eV, and the threshold condition is considerably affected by the violation effect for $E > 10^{13} \xi_{\gamma} ^{-1/3}$ eV, which can be much lower than the Planck energy scale. Let us examine the above speculation by calculating in our rest frame the conservation of momentum of the reaction. By putting $c$ = 1 hereafter, the momentum $E - \varepsilon + \xi _{\gamma} E^2 /(2E_0)$ of the initial state (we assume for simplicity, but not to lose the generality, $\lq$head-on' collision of $\gamma$-ray and soft photon) must be equal to $p_1 cos \alpha + p_2 cos \beta$ of the final state, where ($E_1$, $\vec p_1$) and ($E_2$, $\vec p_2$) is energy and momentum of electron and positron and $E + \varepsilon = E_1 + E_2$. The $\alpha$ and $\beta$ are the angles $\vec p_1$ and $\vec p_2$ make against the direction of the initial gamma ray and are in the order of $m / E_i$ ($i = 1$ and 2). By using the approximation $p_i = E_i - m^2/(2E_i)$ and the condition on the transverse momenta $\alpha p_1 = \beta p_2$, we obtain \begin{equation} -2 \varepsilon + {\xi _{\gamma} E^2 \over 2 E_0} = - {E + \varepsilon \over 2} \cdot ( {m^2 \over E_1 E_2} + \alpha \beta ). \end{equation} We use a parameter $x$ to replace $E_1$ and $E_2$ by $E_{1, \, 2} = (1 \pm x) (E + \varepsilon )/2$ with the condition $| x | \le 1 - 2m /(E + \varepsilon )$. The equation (2) is then written as \begin{equation} \varepsilon \approx {m^2 \over E} \cdot {1 \over 1 - x^2 } + {\alpha \beta E \over 4} + \xi _{\gamma} \cdot {E^2 \over 4 E_0}. \end{equation} \noindent The first and second terms are of the same order of magnitude, the latter term approaching zero at the threshold of the reaction. With increasing $E$, softer photons can serve as the {\it target} photons, shortening the absorption mean free path. This conventional view is altered as $E$ further increases, by the presence of the term $\xi _{\gamma} E^2 / (4 E_0)$. However, electron and positron may also obey a modified relation similar to (1). Let us put, similarly to (1), the square of the modified momentum $q$ to be equal to \begin{equation} q^2 = p^2 + \xi _e {E^3 \over E_0} = E^2 - m^2 + \xi _e {E^3 \over E_0}. \end{equation} \noindent We grant $|\xi _e | \ne 1$ for allowing electrons to have a different degree of the invariance violation. By replacing $p$ by $q$, a relation similar to (3), \begin{equation} \varepsilon = {m^2 \over E} \cdot {1 \over 1 - x^2 } + {\alpha \beta E \over 4} + {E^2 \over 4 E_0} (\xi _{\gamma} - \xi _e \cdot {1 + x^2 \over 2 }), \end{equation} \noindent is obtained. Since energy $E$ is divided almost equally into two particles (when $x \approx 0$) and the anomalous term has $E^2$ dependence, the factor 2 appears in the denominator of $\xi _e (1 + x^2)/2$, thus reducing the effect in the final state. The fact might seem to suggest that the single $\lq$leading' particle taking almost all the total energy ($|x| \approx 1$) is capable of eliminating the anomalous effect. However, the substitution of $|x| = 1 - 2m /(E + \varepsilon )$ into (5) increases the energy $\varepsilon$ to $m \sim 1$ MeV which is much higher than $m^2 /E$, and then leads to a very large travelling distance, resulting in no less anomalous consequence. Only if $\xi _e$ is tuned to be $2\xi _{\gamma}$ for unknown reason, we can recover the ordinary relation for $\varepsilon$. \section{Comparison with observations} Energy $\varepsilon$ of the {\it target} photon is plotted against $\gamma$-ray energy $E$ in Fig.~1. A minimum value of $\varepsilon$ is taken, if $\xi = 2 \xi _{\gamma} - \xi_e > 0$, at the critical energy $E_c = (4m^2 E_0 /\xi )^{1/3} \approx 10^{13}$ eV for $\xi = 1$. The minimum value is about 0.03 eV and $\varepsilon$ must increase to about 30 eV at $E = 10^{15}$ eV. Thus, CMB of $\sim 10^{-3}$ eV is never given a chance of contributing to the pair creation process. The travelling length has the shortest value of $\sim$10 Mpc at $E = E_c$, and increases again for $E > E_c$. Since $\sim 30$eV photons are very sparse, $\gamma$-rays would travel through distances $>$ 1 Gpc. When $\xi <0$, $\varepsilon$ must rapidly decrease as $E$ approaches $E_c$. Before taking negative value, CMB photons become available as {\it target}. The energetic $\gamma$-rays are then efficiently absorbed with a mean free path of $\sim 10$ kpc, which is shorter by orders of magnitude than the ordinary case. \placefigure{fig1} Among the TeV $\gamma$-ray sources are two active galaxies, Mrk 421 and Mrk 501 (Punch et al. 1992; Aharoninan et al. 1997; Catanese et al. 1997). The objects are at distances of $\sim$ 100 Mpc, and the detection is consistent with the absorption mean free path determined by infrared intensity in the extra-galactic space. The energy spectrum of the $\gamma$-rays is found to extend up to 10 - 20 TeV. If the spectrum continues to extend to even higher energies, $\xi < 0$ would be excluded. Because, if $\xi < 0$, a very sharp cut-off at the common energy $E_c$ is expected in the energy spectrum for any extra-galactic objects. The negative $\varepsilon$ for $E > E_c$ in the case of $\xi < 0$ implies that $\gamma$-rays does not need {\it target} photons for pair creation. The threshold of $\gamma$-ray decaying into an electron-positron pair (Coleman and Glashow 1997) is given by $|E_c|$. By putting $|E_c| = -E_c = (4m^2 E_0/(-\xi ))^{1/3} > 20$ TeV, we obtain a limit on $\xi$ to be $\xi > -1.3$ (by putting $E_0 = 1 \times 10^{28}$ eV). On the other hand, when $\xi$ is positive and $\sim 1$, the travelling distance turns over to increase above about 10 TeV and to become larger than 1 Gpc for the photons around 100 TeV. Thus, 100 TeV photons would arrive from extra-galactic sources, implying that evidence of detection would be strongly in favour of the $\xi >0$ case. Extra-galactic diffuse $\gamma$-rays are most likely from distant unresolved active galaxies. The energy spectrum is observed by Sreekumar et al. 1997 to extend to about 50 GeV with a hard power index of $\sim -2.3$. The spectrum is expected to become steeper at higher energies due to the intenser infrared photons and then CMB at longer wavelengths. However, in the case of $\xi >0$, the energy spectrum may have a {\it turn-over} of showing a harder slope above $E = E_c$. Electrons lose their energy quickly and have difficulty in producing such high energy $\gamma$-rays. However, cosmic ray protons are known to have energy spectrum extending to $10^{20}$ eV, and are able to produce $\gamma$-rays as high as or larger than 100 TeV. The increase of travelling distance suggests more contribution from more distant sources and thus we may expect considerably intense, diffuse $\gamma$-rays in the higher energy region. From the Crab nebula, 50 TeV photons have been detected (Tanimori et al. 1997) by using imaging \v Cerenkov telescope at large zenith angles. Detection of 100 TeV $\gamma$-rays is not unrealistic by the extensive application of such methods. \section{Case of cosmic ray protons} Our phenomenological consideration can be extended to other processes of soft photons acting as the reaction target. Extremely high energy cosmic ray protons ({\it e.g.} Hill and Schramm 1985; Axford 1994; O'Halloran et al. 1998; Takeda et al. 1998) are believed to lose energy by pion production in collision with CMB. The number density of CMB and the photo-pion cross section determine the collision mean free path to be $\sim 10$ Mpc. The energy spectrum thus suffers from the GZK cutoff (Greisen 1966; Zatsepin and Kuzmin 1966) at about $10^{20}$ eV, if the protons are from the distances farther than the mean free path. However, the relation (4) assumed for proton and pion alters the threshold of the energy $\varepsilon$ of {\it target} photons from $\mu M /(2E)$ to \begin{equation} \varepsilon = {\mu M \over 2E} + \Xi \cdot {E^2 \over 4 E_0} , \ \ {\rm where } \ \ \Xi = \xi _p - \xi _p ({M \over M + \mu })^2 - \xi _{\pi} ({\mu \over M + \mu })^2. \end{equation} \noindent The energy $E$ is for the initial proton, $M$ and $\mu$ are the proton and pion mass, and $\xi _p$ and $\xi _{\pi}$ for protons and pions, respectively. We have put the parameter $x$ to be $(M -\mu) / (M+\mu)$, corresponding to the case in which the ratio of proton to pion energy in the final state is $(1+x)/(1-x) = M / \mu$ near the threshold of the process. If $\Xi$ is positive and not very far from the order of one, the second term becomes comparable with $ \mu M /(2E)$ at the critical energy $E_{p\pi} = ( M \mu E_0 /\Xi)^{1/3} \approx 10^{15}$ eV. At $E \approx 10^{20}$ eV, the energy $\varepsilon$ needs to be, as dominated by $\Xi E^2 /(4E_0)$, as large as $10^{12}$ eV, and CMB can not play the role of {\it target} photons. Cosmic ray protons at $\sim 10^{20}$ eV can then travel without the attenuation from cosmological distances farther than $\sim 10$ Mpc. As a result, the energy spectrum of cosmic rays would extend beyond $10^{20}$ eV, provided that the acceleration to such high energies takes place in celestial objects. There is speculation by Vietri 1995 and Waxman 1995 that GRB ($\gamma$-ray bursts) provides the origin of the highest energy cosmic rays. If so, GRBs at cosmological distances would produce a cosmic ray spectrum without GZK cutoff. If $\Xi < 0$, $\varepsilon$ can be $\sim 10^{-3}$ eV when $E$ approaches $|E_{p\pi}| = - E_{p\pi} \approx 10^{15}$ eV, implying that the {\it target} photons of CMB gives a characteristic attenuation length of $\sim$10 Mpc to $10^{15}$ eV cosmic rays. For $E > |E_{p\pi}|$, the kinematical condition does not require the {\it target} photon for the pion production to occur and the processes of $p \rightarrow p + \pi^0$ and $p \rightarrow n + \pi^+$ will cause protons to rapidly lose energy. The energy loss due to the spontaneous production of pion is in contradiction with the evidence of cosmic ray protons up to $10^{20}$ eV. Thus, we limit $|E_{p\pi}|$ longer than $10^{20}$ eV, and obtain a bound of $\Xi > - 10^{-17}$. \section{Discussions} Collision processes of high energy radiations with soft photons enable us to test the consequences of violation of Lorentz invariance. In particular, the effect of quantum gravity which Amelino-Camelia et al. (1998) suggest can be made detectable by the reaction process in the energy region much lower than the Planck energy scale, but in the sufficiently high energy region to which the validity of Lorentz transformation has not been examined yet by accelerator experiments. The violation suggests an interesting possibility of increasing the distance we can reach by observing high energy $\gamma$-rays and cosmic rays. Absence of the GZK cutoff would not necessarily mean nearby origin of the extremely high energy cosmic rays. Photons of energies larger than 100 TeV from extra-galactic objects would provide a clear test because it is hard to explain by other mechanisms than the violation of invariance. It seems impossible, once we admit the relation (1) for photons, to totally eliminate the anomalous effect, without a specially designed tuning among the invariance violating terms for various particles of different masses. Other ways have been proposed to alter the conventional view on the relativity (Gonzares-Mestres 1997; Sato 1998). Absence of GZK cutoff can be explained, according to Sato 1998, without abandoning the Lorenz invariance, but by giving a special status to $\lq\lq$the universe frame'' in which CMB is isotropic. Protons are expected to lose energy via vacuum \v Cerenkov radiation (Coleman and Glashow 1997), when $\xi _{\gamma} -\xi _p >0$. The threshold energy, $E_{th} = (M^2 E_0 / (\xi _{\gamma} - \xi _p))^{1/3}$, of the radiation is calculated from equating the speed of photon to that of proton. The detection of the highest energy $10^{20}$ eV of cosmic rays sets a constraint $E_ {th} > 10^{20}$ eV and limits the $\xi$-parameters to be $\xi _{\gamma} -\xi _p < 10^{-15}$ or rather $\xi _{\gamma} \le \xi _p$ for practical use. The \v Cerenkov radiation from electron yields a less tight constraint because of the dependence of $\xi$ on $E_{th}$, {\it i.e.} $\xi = \xi_{\gamma} - \xi_e \propto E_{th}^{-3}$. The highest energy $2 \times 10^{12}$ eV of cosmic ray electrons ever detected (Nishiumura et al. 1997) must be less than $E_{th}$ and then we obtain $\xi _{\gamma} -\xi _e < 10^3$. If we choose, for the bound on $E_{th}$, $\sim 100$ TeV electrons which are responsible for TeV $\gamma$-rays and which are considered to extend over a spatial scale of at least $\sim$ 1 pc, the constraint is made somewhat tighter as $\xi _{\gamma} -\xi _e < 10^{-2}$. We may argue about allowed region of $(\xi _{\gamma}, \xi_e)$ values by combining the condition $-1.3 < 2\xi _{\gamma} - \xi _e$ from the decay process $\gamma \rightarrow e^+ + e^-$. The region is, to some extent, in disfavor with negative values. The time structure of the outbursts of TeV $\gamma$-rays from Mrk~421 and Mrk~501 sets constraint on $\xi _{\gamma}$. The delay time of 10 TeV $\gamma$-rays relative to photons of lower energies is $(d/c)\xi_{\gamma} (10^{13} {\rm eV} /E_0) \approx 10 \xi_{\gamma} \, s$ for distance $d=100$ Mpc. Duration of the outbursts which has been observed typically as a few hours is considered to be longer than the delay time, providing a constraint of $\xi_{\gamma} < 10^3$. We need to know the outburst structure in time scales as short as $\sim 10 s$ to constrain $|\xi|$ less than $\sim 1$. In order to restrict the $|\xi|$ parameter more tightly, It is very useful to observe photons of higher energies from more distant objects. However, we can not expect that such $\gamma$-rays survive against conversion into an electron-positron pair and reach us, unless we premise that the Lorenz invariance is violated. We may argue that the extensive air shower events of $\sim 10^{20}$ eV are in truth not due to proton. Such a claim is relevant to $\lq$correct' understanding of the complex phenomena of the cascade shower process, which still leaves a room for a speculation of $\lq$anomalous interaction' of cosmic rays. Thus, it is interesting to indicate that the modified relation of energy and momentum can affect detection of the high energy radiations. Observation of $\gamma$-rays for example is based on the pair creation process in the detector material (satellite experiments) or in the atmosphere (ground-based observation for TeV photons and cosmic rays). The recoil momentum $\delta$ of atomic nucleus in the material is calculated as twice the right hand side of the expression (5). The ordinary recoil momentum is $\delta _0 \approx m^2 E/(2E_1 E_2) = 0.1$ eV/c at $E \approx 10^{13}$ eV, which corresponds to quite a small energy transfer to the recoiled atomic nucleus of a large mass. When $\xi _{\gamma} =1$ and $E > E_c$, $\delta$ must be larger than $\delta _0$ but remains reasonably small so that the process would not be largely affected. When $\xi _{\gamma} = - 1$ and $E > E_c$, $\delta$ appears negative, the recoil being in the opposite direction to the initial $\gamma$-ray. There are generally several processes which compete the pair creation and for which the invariance violation is of different strengths. The consequence may appear as observation of an anomalous behaviour of very high energy radiations. \acknowledgments The author thanks the anonymous referee for the comments and advice that have brought the revised form and content of the present paper.
\section{Introduction} \label{intro} \footnotetext[1]{Submitted to PRE Feb. 1999.} Collections of inelastically colliding macroscopic particles, when driven sufficiently by an external force, exhibit fluid like behavior such as flow~\cite{campbell90,jaeger96} and instabilities~\cite{melo94,debruyn98}. Although only a small number of particles ($10^2-10^6$) are typically involved in a flow, continuum methods are popular tools in modeling, because a century of fluid dynamics experience can be brought to bear on the problem. With continuum equations velocity profiles and transfer rates can be calculated, and stability analyses performed. Often, plausible continuum equations are simply posited, but these equations typically contain unmeasurable free parameters. A more rigorous approach derives continuum equations from the kinetic theory of dissipative gases~\cite{lun84,jenkins85,jenkins85a}. In principle, a closed system of partial differential equations results, analogous to the Navier-Stokes equations, with all transport coefficients given. However, the continuum equations for granular media do not share the stature of their molecular fluid analogs. For both grains and molecules, the general form of the equations is not in question, but the constitutive equations relating fluxes of momentum and energy to gradients may differ from the simple cases of Newton's viscosity law and Fourier's heat law. The transport coefficients of liquids are not routinely calculated from kinetic theory, but are measured experimentally; such measurements have not been carried out for granular media, where the granular temperature (the kinetic energy associated with the fluctuational velocities of particles) is difficult to control and difficult to measure. In addition, the small number of particles and the dissipative nature of the medium have led some researchers to the conclusion that continuum approaches are doomed to failure~\cite{kadanoff97}. As a result, the central question remains open: {\emph{Can continuum equations, derived from kinetic theory and supplemented by measurements, model rapid granular flows to the same level that Navier-Stokes equations model the flows of liquids?}} Molecular dynamics simulations will play a crucial role in answering this question, since the data they provide can be used to quantitatively test the assumptions and the results of kinetic theory. The main assumptions, which we will elucidate in Sec.~\ref{kinetictheory}, are: \begin{enumerate} \item Single particle distribution functions are nearly Boltzmann, \item Molecular chaos --- Particle velocities are uncorrelated, and \item Particle positions are correlated in accord with Eq.~(\ref{CandS}), the Carnahan and Starling relation for elastic particles. \end{enumerate} The main results, also discussed further in Sec.~\ref{kinetictheory}, are: \begin{enumerate} \item The equation of state (Eq.~(\ref{state})), \item The constitutive relations (Newton's stress law and Fourier's heat law --- Eqs.~(\ref{newtonslaw}) and~(\ref{fourierslaw})), and \item The values of the shear viscosity $\mu$, the thermal conductivity $\kappa$, and the loss rate of granular temperature due to inelastic collisions, $\gamma$ (Eqs. ~(\ref{viscosityenskog}),~(\ref{conductivityenskog}), and ~(\ref{gamma0eq})). \end {enumerate} We will use molecular dynamics simulations to test these points, and to measure the transport coefficients to be used in continuum analyses of granular flows. We have found that the simulations quantitatively reproduce experiments on standing waves in oscillated granular media, producing the correct wave patterns and wavelengths~\cite{bizon98} and secondary instabilities~\cite{debruyn98}. With the numerical simulation, we not only have access to experimentally unmeasurable quantities~\cite{bizon98a}, but we also can study systems that are not experimentally realizable, allowing us to test both the assumptions and results of the granular kinetic theory. Once the constitutive relations have been tested and possibly enhanced through the use of particle simulations, direct comparison between continuum theory and laboratory experiment becomes possible. Laboratory experiments of granular kinetic theory have not yet proceeded beyond measuring the single particle velocity distribution function~\cite{olafsen98,delour98,oger96}. Simulations that make contact with kinetic theory have focussed mainly on the homogeneous cooling state, a time-dependent state that eventually becomes spatially inhomogeneous. In those simulations, long-range velocity correlations develop~\cite{orza98}, and molecular chaos no longer holds. A more sophisticated form of kinetic theory, ring kinetic theory, in which correlations in particle velocity are accounted for, is required for a full description~\cite{vannoije98}. Finally, a number of simulations have been performed on one-dimensional granular media in contact with a heat bath~\cite{williams96,swift98,puglisi98} to produce a steady state. In one-dimension (1D), spatial~\cite{williams96} and velocity~\cite{swift98} correlations can develop, and the single particle velocity distributions may be nongaussian~\cite{puglisi98}. The present work extends these 1D steady state and 2D decaying simulations by examining 2D steady states with the goal of quantitatively testing the assumptions and results of granular kinetic theory. In Section~\ref{sims} we describe the molecular dynamics simulations and the forcing methods. Section~\ref{homogeneous} discusses simulations of granular media in which the heat bath is spatially homogeneous. In that section, we will check the assumptions about the nature of the single particle distribution function and the correlations of position and velocity, as well as measuring the equation of state and $\gamma$. In Section~\ref{inhomogeneous} we turn to our main purpose: allowing the heat bath to vary spatially so that steady state inhomogeneous states may be induced and described. From these simulations we can study the constitutive relations, the shear viscosity, and the thermal conductivity. Section~\ref{conclusion} contains concluding remarks. \section{Kinetic Theory} \label{kinetictheory} We begin with a brief review of the kinetic theory of granular media, which differs only slightly from the kinetic theory of elastic particles as presented in textbooks such as~\cite{chapman}. The number of particles in a volume and velocity element $d{\bf r}d{\bf v}$ centered at position ${\bf r}$ and velocity ${\bf v}$ is given by ${f^{(1)}({\bf r},{\bf v},t)} d{\bf r}d{\bf v}$; ${f^{(1)}({\bf r},{\bf v},t)}$ is called the single particle distribution function. Continuum quantities are given as averages over ${f^{(1)}({\bf r},{\bf v},t)}$. In particular, the number density, average velocity and granular temperature are defined respectively as \begin{eqnarray} n({\bf r},t) &\equiv& \int {f^{(1)}({\bf r},{\bf v},t)} d{\bf v},\\ {\bf v}_0({\bf r},t) &\equiv& {{1}\over{n}} \int {f^{(1)}({\bf r},{\bf v},t)} {\bf v} d{\bf v},\\ T({\bf r},t) &\equiv& {{1}\over{nD}} \int {f^{(1)}({\bf r},{\bf v},t)} ({\bf v}-{\bf v}_0)^2 d{\bf v}, \end{eqnarray} where $D$ is the number of dimensions. Note that the granular temperature $T$ is not the thermodynamic temperature of the particles due to the random motions of their molecules, but the analogous kinetic energy due to the random motions of the macroscopic particles themselves. The evolution of ${f^{(1)}({\bf r},{\bf v},t)}$ depends on the joint probability distribution, ${f^{(2)}({\bf r_1},{\bf v_1},{\bf r_2},{\bf v_2},t)}$; collisions of two particles change the single particle distribution. As in molecular kinetic theory, Boltzmann's assumption of molecular chaos is made, {\it i.e.,}\ velocities are assumed to be uncorrelated, although for sufficiently dense media, positional correlations are allowed: \begin{eqnarray} \lefteqn{f^{(2)}({\bf r}_1,{\bf v}_1,{\bf r}_2,{\bf v}_2,t) = } \nonumber \\ & & g(\sigma,\nu) f^{(1)}({\bf r}_2-\sigma {\bf\hat{k}},{\bf v}_1,t) f^{(1)}({\bf r}_2,{\bf v}_2,t), \label{gchaos} \end{eqnarray} where $\sigma$ is the particle radius, ${\bf\hat{k}}$ is a unit vector pointing from the center of particle 1 to the center of particle 2, and $\nu ={{1}\over{4}} n \pi \sigma^2$ is the solid fraction in 2D. The positional correlations are accounted for through $g(r,\nu)$, the radial distribution function, which is defined as the probability of having a pair of particles whose relative distance lies in the interval $r,r+dr$, normalized by the probability for an ideal gas. This function, evaluated at the point of contact $r=\sigma$, gives the increase in the probability of collisions due to dense gas (excluded volume) corrections. For elastic hard disks, spatial correlations are described by the formula of Carnahan and Starling~\cite{carnahan69}, \begin{equation} G_{CS}(\nu) = {{\nu(16-7\nu)}\over{16(1-\nu)^2}}, \label{CandS} \end{equation} where \begin{equation} G(\nu) \equiv \nu g(\sigma,\nu). \label{Gg} \end{equation} Equation~(\ref{CandS}) works well for elastic particles with solid fractions below 0.675, where a phase transition takes place~\cite{alder62}, and is often used in modeling granular media~\cite{jenkins85a}. Equation~(\ref{Gg}) is the definition of $G$ in terms of the (unknown) radial distribution function $g(r,\nu)$, evaluated at $r=\sigma$, while Eq.~(\ref{CandS}) is a particular model for $G$, denoted by the subscript $CS$. An unforced collection of molecules approaches a Boltzmann distribution, \begin{equation} {f^{(1)}({\bf r},{\bf v},t)}=n (2 \pi T)^{-D/2} e^{-C^2 / 2T}, \end{equation} where $C \equiv |{\bf v}-{\bf v}_0|$. Away from equilibrium, the local distribution for elastic particles is nearly Boltzmann. Granular media dissipate energy with each collision, so that the equilibrium state of an unforced granular medium is that of no relative motion. For grains, the single particle distribution function is simply assumed to be nearly a Boltzmann distribution. With these assumptions, and the additional assumption that the coefficient of restitution $e$ is only slightly less than 1 (particles are only slightly inelastic), equations for the continuum mass, momentum, and energy can be derived for disks in two dimensions~\cite{jenkins85,jenkins85a}: \begin{eqnarray} {{\partial n}\over{\partial t}} + {\bf \nabla} \cdot (n{\bf v_o})&=&0\\ n{{\partial {\bf v_o}}\over{\partial t}} + n{\bf v_o}\cdot{\bf \nabla v_o} &=& -{\bf \nabla \cdot \underline P}\\ n{{\partial T}\over{\partial t}} + n{\bf v_o}\cdot{\bf \nabla} T &=& -{\bf \nabla} \cdot {\bf q} - {\bf \underline P}:{\bf \underline E}-\gamma, \ \label{balance} \end{eqnarray} where $E_{ij}={{1}\over{2}}(\partial_i v_{oj} + \partial_j v_{oi})$ are the elements of the symmetrized velocity gradient tensor ${\bf \underline E}$. The constitutive relations for the pressure tensor ${\bf \underline P}$ and heat flux ${\bf q}$ are \begin{equation} {\bf \underline P} = (P - 2 \lambda {\rm Tr}{\bf \underline E}){\bf \underline I} - 2 \mu ({\bf \underline E} - ({\rm Tr}{\bf \underline E}){\bf \underline I}) \label{newtonslaw} \end{equation} and \begin{equation} {\bf q} = - \kappa {\bf \nabla} T, \label{fourierslaw} \end{equation} where Tr denotes trace and ${\bf \underline I}$ is the unit tensor. The 2D equations close~\cite{jenkins85a} with the equation of state, which is the ideal gas equation of state with a term that includes dense gas and inelastic effects, \begin{equation} P = (4/\pi\sigma^2) \nu T [1 + (1+e) G(\nu)] \label{state}, \end{equation} and the predicted values, denoted with a subscript $0$, for the bulk viscosity, $\lambda$, \begin{equation} \lambda_0 = {{8 \nu G(\nu)}\over{\pi \sigma}} \sqrt{{T}\over{\pi}} \end{equation} the shear viscosity, $\mu$, \begin{equation} \mu_0 = {{\nu }\over{2 \sigma}} [{{1}\over{G(\nu)}} + 2 + (1+{{8}\over{\pi}})G(\nu)] \sqrt{{T}\over{\pi}}, \label{viscosityenskog} \end{equation} the thermal conductivity, $\kappa$, \begin{equation} \kappa_0 = {{2 \nu }\over{\sigma}} [{{1}\over{G(\nu)}} + 3 + ({{9}\over{4}}+{{4}\over{\pi}})G(\nu)] \sqrt{{T}\over{\pi}} , \label{conductivityenskog} \end{equation} and the temperature loss rate per unit volume, $\gamma$ \begin{equation} \gamma_0={{16 \nu G(\nu)}\over{\sigma^3}}(1-e^2)\left({{T}\over{\pi}}\right) ^{3/2}. \label{gamma0eq} \end{equation} Because of the assumption of near elasticity, the coefficient of restitution enters only in the equation of state and in the expression for the temperature loss due to inelastic collisions. To this order, the thermal conductivity and shear viscosity are the same as those given by the Enskog procedure for elastic disks~\cite{gass70}. \section{Driven Granular Media Simulations} \label{sims} We perform event-driven simulations~\cite{lubachevsky91,marin93} of a granular gas in a two-dimensional periodic box of side length $L = 52.6 \sigma$, in contact with a thermal bath that stochastically heats particles throughout the volume. Between collisions, particles travel freely. In our model the collisions are instantaneous and binary; they conserve momentum and dissipate energy. Particles are assumed to be frictionless; particle friction can be incorporated into kinetic theories~\cite{lun87,lun91}, and was included in the simulations of oscillated granular media,~\cite{bizon98,bizon98a,debruyn98}, but introduces complications that we wish to avoid for this study. When particles collide, new velocities are calculated by reversing the component of the relative particle velocity along the line joining particle centers and multiplying it by the coefficient of restitution $e$, which is between 0 and 1. If $e$ is independent of collision velocity, a finite time singularity can occur in the collision frequency, a phenomenon known as inelastic collapse~\cite{mcnamara92,mcnamara94}. To avoid this simulation-ending occurrence in a convenient and natural way, we allow $e$ to vary as \begin{equation} e(v_n) = \left \{\begin{array}{cc} 1 - B v_n ^ {\beta} &, v_n < v_a \\ \epsilon &, v_n > v_a \end{array} \right. , \label{restform} \end{equation} where $v_n$ is the component of relative velocity along the line joining particle centers (normal to the contact surface), $B = (1-\epsilon)(v_a)^{-\beta}$, $\beta=3/4$ and $\epsilon$ is a constant, chosen to be $0.7$. In simulations of oscillated granular media, the results are not sensitively dependent on $v_a$ or $\beta$, and $\epsilon=0.7$ produces good agreement with experiment~\cite{debruyn98,bizon98,bizon98a}. In addition to forestalling collapse, variation of $e$ allows us to further probe granular kinetic theory, which assumes that $e \approx 1$. As $T$ varies, the relative collision velocity will vary; hence so will $e$, and the relative importance of that assumption can be gauged. At a given temperature, a distribution of collision velocities and a corresponding distribution of coefficients of restitution occur. Varying $T$ varies not only the average value of $e$, but also the amount of deviation around that average. The variation in $e$ gives rise to a velocity scale that is not present in the elastic case. All quantities given below are nondimensionalized with the particle diameter $\sigma$ and the crossover velocity $v_a$ at which the coefficient of restitution becomes a constant. In particular, the granular temperatures all scale with $v_a^2$. Because inelastic collisions remove energy from the system, we must constantly add energy to achieve any sort of steady state. The situation is opposite that in simulations on nonequilibrium systems of elastic particles, where the constant energy input from the driving must be removed through an artificial means~\cite{evans}. Stochastic heating is performed in one of three ways: A) \emph {White Noise} --- random kicks $\delta {\bf v}$ are added to particles' velocities, B) \emph{Random Accelerations} --- particles accelerate between collisions, and C) \emph {Boltzmann Bath} --- particle velocities are obliterated and and replaced with velocities chosen from a Gaussian distribution. We discuss the motivation and implementation of each in turn. \subsection{White Noise} Williams and Mackintosh~\cite{williams96} introduced white noise as a thermal bath for dissipative granular media. In their model, a random velocity is added to each particle's velocity during each time step $\Delta t$. The velocities added to each particle are not correlated with one another, nor are they correlated with the velocities added in the previous $\Delta t$. This model has the advantage that the equation of motion for particles between collisions may be written down as the Langevin equation \begin{equation} {{d^2x_i}\over{dt^2}} = \zeta_i, \label{Langevin} \end{equation} where $x_i$ is the position of the $i$-th particle and $\zeta_i$ is a Gaussian white noise term, {\it i.e.,} ~$\langle\zeta_i(t)\zeta_j(t')\rangle = 2 F \delta_{ij} \delta(t-t')$. The fact that the heating can be analytically expressed makes its inclusion into kinetic theory possible. This forcing is straightforward to include in our simulations. Rather than adding the random kicks to particles all at once, we kick $2r_k$ randomly chosen particles every time there is a collision. The number $r_k$, then, represents the ratio between the rate of kicks and the rate of collisions. If the kicks are totally random, the center of mass momentum of the system will fluctuate, but we desire that the heat bath can only change the fluctuational velocity, not the mean. To ensure that the mean velocity remains fixed, we apply random kicks to $r_k$ particles: \begin{equation} {\bf v}_i \rightarrow {\bf v}_i + |\delta {\bf v}| \hat{\bf r}_i, 1\le i\le r_k. \end{equation} The kicks themselves are all of the same size, $|\delta \bf v|$, but the directions $\hat{\bf r}_i$ are randomly chosen. Then, kicks in the opposite directions are applied to another $r_k$ randomly chosen particles: \begin{equation} {\bf v}_i \rightarrow {\bf v}_i - |\delta {\bf v}| \hat{\bf r}_{i-r_k}, r_k<i\le 2 r_k. \end{equation} Because each kick requires recalculation of the kicked particle's collision list, we want to minimize the kick frequency. Empirically, we find that our results are independent of $r_k$ for $r_k \ge 1$ and $r_k(\delta{\bf v})^2$ constant. This is not surprising, since the average length of $r_k$ randomly oriented kicks of length $\delta\bf v$ is $\sqrt{r_k}\delta v$, and only the total kick between collisions matters when collisions occur. For this reason, we perform most of our simulations with $r_k = 1$, where the collision rate equals the kick rate. \subsection{Random Accelerations: The Air Table Model} \label{accelerated} While white noise forcing has the advantage that it can be incorporated into the kinetic theory relatively easily, it has the disadvantage that it does not model any particular real system. In most experiments, energy is added to the granular media through a boundary, causing gradients in the energy perpendicular to that boundary. For a vertically oscillating layer in a gravitational field~\cite{melo94}, the situation is even more severe; although one might like to imagine the oscillating wall as thermal, providing stochastic kicks to the layer, the interactions of the layer with the plate are strongly correlated to the plate's motion. The plate introduces both spatial and temporal gradients. One experimental system is capable of producing homogeneous steady states, specifically, a collection of pucks on an air table~\cite{oger96,ippolito95}. Although the purpose of the air is to levitate the pucks, it also drives their horizontal motion. Either due to inhomogeneities in the air flow or because the pucks are not perfectly parallel to the surface of the table, the pucks accelerate uniformly from one collision to another~\cite{oger96}. This driving counters the loss of energy due to inelastic collisions and produces a steady state. We model this air table driving by allowing each particle to move under a uniform acceleration: \begin{equation} {\bf a}_i = a_0 \hat{\bf r}_i \end{equation} The magnitudes of all particle accelerations, $a_0$, are the same, but the directions, $\hat{\bf r}_i$, are randomly and uniformly chosen. The direction of a particle's acceleration changes stochastically. When a collision occurs, $2r_k$ particles are given new $\hat{\bf r}_i$. In order to conserve total momentum, we hold the total acceleration of the particles at zero by giving exactly opposite accelerations to pairs of particles. Initially, each particle is paired with another, and these are given opposite accelerations. Later, when one particle is chosen and its acceleration randomized, its partner particle is also given a new acceleration, opposite to the first particle's. Experiments suggest that particles accelerate uniformly from collision to collision, with most of the changes in acceleration happening at collisions~\cite{oger96}. Therefore, $r_k=1$ is probably a relatively good model for the air table experiments. As $r_k$ increases, the particles feel a constant acceleration over a small temporal range. In the limit that $r_k\rightarrow\infty$, the model is the same as the white noise model, which is rate independent. In practice, we observe that the distribution of collision velocities for the accelerated particle model approaches that for the white noise model at $r_k \approx 8$. Variation of coefficient of restitution with normal collision velocity is critical for simulations with the accelerated forcing, since inelastic collapse-like collision sequences are more prevalent. After a collision, particles move away from one another. Without relative acceleration, they must collide with other particles to reverse their velocity if they are to recollide. Particles with relative acceleration, however, may collide and recollide without striking another particle in the intervening time, just as a ball bounces repeatedly on the ground. On average, the relative acceleration is likely to change during this interval, but that is not guaranteed for any given pair. Further, since accelerations are changed when collisions occur, the rate of acceleration fluctuation is dependent on the collision rate in the entire medium. A given pair of particles can hijack the collision sequence of the gas, rapidly recolliding with one another. If the coefficients of restitution are constant, this scenario will produce collapse. By allowing the collisions to become more elastic for decreasing relative velocities, however, collapse is prevented; eventually, the relative acceleration will change, and the particles will move apart. \subsection{The Boltzmann Bath} Finally, we introduce a heat bath that approximates the assumption of molecular chaos. Molecular chaos assumes that the velocities of colliding particles are uncorrelated; particles collide, and before they collide with one another again, they collide with a large number of other particles, losing the memory of the initial condition. A strong heat bath can perform the same function if it replaces the particle velocities with new velocities chosen from a given distribution. If the heat bath interacts with the particles often enough, molecular chaos will be guaranteed. While this bath is wholly unphysical, it produces a situation in which the kinetic theory is expected to apply exactly; it is a useful check on calculations, and helps to elucidate the role of velocity correlations. Implementation of the Boltzmann bath is simple. When two particles collide, $2r_k$ particles are randomly chosen and given velocities chosen from a Boltzmann distribution with a specified temperature. \section{Homogeneous Forcing, Correlations, and the Equation of State} \label{homogeneous} With any of the thermal baths just described, we can set up inhomogeneous states by varying the strength or rate of forcing over space. However, in the simplest case we force homogeneously. The simulations are performed for a variety of solid fractions, forcing rates, and forcing strengths for the three thermal baths. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{evsT.ps}} \smallskip \caption[Average coefficient of restitution versus temperature] {Average coefficient of restitution vs T for $\nu=0.5$, for the three forcing methods. $+$: White noise forcing, $\diamond$: Accelerated forcing, $*$: Boltzmann Bath. As discussed in the text, $T$ is nondimensionalized with $v_a^2$.} \label{restitution} \end{figure} The average coefficients of restitution for a number of runs at different temperatures are shown in Fig.~\ref{restitution}. Note that at high temperatures, the average coefficient of restitution begins to rise for the accelerated forcing. This is due to the inelastic collapse-like collision sequence described earlier for this forcing. As the magnitude of acceleration increases to produce higher temperatures, these multiple collisions become more and more important, leading to a large number of collisions with very low velocities, and so a high average coefficient of restitution. In the state produced by homogeneous forcing, we can measure the single particle distribution function, the temperature produced by the forcing, the pressure, the radial distribution function, velocity correlations, and loss rates. These quantities can be compared to the corresponding quantities for elastic simulations and to their assumed or calculated values from kinetic theory. \subsection{Single Particle Velocity Distributions} The lowest order approximation to $f$ in the kinetic theories is usually chosen to be a Boltzmann distribution, the form of $f$ for an undriven elastic gas. A driven inelastic gas, although it approaches a steady state, is by no means guaranteed to act like an undriven elastic gas. Nevertheless, the single particle distribution functions measured from the simulation are all close to Boltzmann distributions, as seen in Fig.~\ref{oneparticledist}. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Pov.ps}} \smallskip \caption[Single particle distribution functions] {Single particle distribution function $\Theta(v^{\prime})$ for (a) and (b) White noise forcing, (c) and (d) Accelerated forcing, and (e) and (f) the Boltzmann bath, all with $r_k=1$. The velocities are scaled with the temperature $T$, so that $v^{\prime} = v/\sqrt{T}$ and $\Theta(v^{\prime}) = Pr(v^{\prime}) \sqrt{T}$, where $Pr(v^{\prime})$ is the probability distribution of $v^{\prime}$. In the left column, the average temperature is approximately 1.05, and the solid fraction is varied ($+$: $\nu=0.1$, $*$: $\nu=0.4$, $\diamond$: $\nu = 0.6$, $\triangle$: $\nu = 0.8$.) In the right column, $\nu$ is fixed at $0.5$ and the temperature is varied; (b) $T= (+) 1.93 \times 10^{-5}, (*)3.13\times 10^{-2},(\diamond)1.06,(\triangle)1067.$ (d) $T = (+) 3.0 \times 10^{-5} ,(*)1.1\times 10^{-2},(\diamond)1.05,(\triangle)256.$ (f) $T = (+) 1.2\times 10^{-5},(*)1.1\times 10^{-2},(\diamond)1.02,(\triangle)102.$ The solid curves are Boltzmann distributions.} \label{oneparticledist} \end{figure} \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Ratio.ps}} \smallskip \caption[Deviations of velocity distribution from Gaussian] {The velocity distribution function $Pr$ of $v^{\prime}=v/\sqrt{T}$ from a simulation with accelerated forcing at $\nu=0.5$ and $T=1.05$, divided by $Pr_{MB}$, a Maxwell-Boltzmann distribution with $T=1.05$. The two curves are for the two velocity components.} \label{ratio} \end{figure} Overall, the accelerated forcing produces the strongest deviations from Maxwellian, and the Boltzmann bath, unsurprisingly, produces the least deviation. In all cases, the deviations become stronger as the density and temperature increase (recall that increasing temperature has the same effect as decreasing the average coefficient of restitution). These deviations tend to flatten the distribution, increasing the probability in the tails and slightly in the peak, and decreasing the probability in between, as displayed in Fig~\ref{ratio}. Similar types of deviations, but much stronger, have been observed in experiments on a dilute, vertically oscillated granular layer~\cite{olafsen98}. \subsection{Equation of State and the Radial Distribution Function} The equation of state,~(\ref{state}), relates the pressure, density and temperature to the coefficient of restitution and $G(\nu)$. The virial theorem of mechanics as applied to hard spheres can be used to calculate the equation of state ~\cite{hirschfelder,rapaportsbook}. \begin{equation} P V = N T + {{\sigma}\over{2 t_m}}\sum_{c} {\bf \hat{k}} \cdot \Delta{\bf v}_i, \label{mdstate} \end{equation} where the sum is over all collisions that occur during the measurement time $t_m $, $\Delta{\bf v}_i$ is the change in the velocity of the i-th particle due to the collision, and $\bf \hat{k}$ is the unit vector pointing from particle center to particle center. In this form, measurement of pressure reduces to measurement of the average particle energy and the average change in the normal velocity at collision; we measure pressure with this method. Using Eq.~(\ref{mdstate}) to measure pressure, and assuming the equation of state (Eq.~(\ref{state})), we produce a measurement of $G(\nu)$, denoted $G_{s}(\nu)$, where the subscript $s$ stands for simulation. This measured value of $G$ will be compared to the Carnahan and Starling value $G_{CS}(\nu)$ from Eq.~(\ref{CandS}). Accurate characterization of $G(\nu)$ is important, because it occurs in the expressions for transport coefficients. \begin{figure}[tb] \epsfxsize=.9\columnwidth \centerline{\epsffile{Gcompare.ps}} \smallskip \caption[Equation of state for inelastic disks] {(a) $G_{s}(\nu)$ for inelastic hard discs driven by $+$: white noise forcing, $\diamond$: accelerations, and $*$: Boltzmann forcing. The solid curve is the Carnahan and Starling relation $G_{CS}(\nu)$, given by~(\protect{\ref{CandS}}). (b) The ratio of $G_s(\nu)$ to $G_{CS}(\nu)$. All runs have $r_k=1$ and $T=1.05$.} \label{Ginelastic} \end{figure} We calculate $G_s(\nu)$ for the three types of forcing as $\nu$ varies. The results are shown in Fig.~\ref{Ginelastic}. For $\nu$ below $\approx 0.675$, where elastic particles undergo a phase transition to an ordered state~\cite{alder62}, the white noise and accelerated runs produce lower $G$ than elastic runs; Boltzmann runs have $G_s(\nu)$ slightly above the elastic values. As the temperature decreases, $e\rightarrow 1$, and the values of $G_s(\nu)$ must approach the elastic values. Therefore, $G_s(\nu)$ must be temperature dependent; this dependence is shown in Fig~\ref{GvsT}, along with the value of $G_s(\nu)$ given by Eq.~(\ref{CandS}). As $T$ decreases, the inelastic $G_s(\nu)$ approaches the elastic $G$, and at high $T$, where $e$ is independent of $T$, $G_s(\nu)$ becomes independent of $T$. As for the single particle distributions, the accelerated forcing shows the greatest deviation from the elastic behavior. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{GvsT.ps}} \smallskip \caption{$G_s(\nu)$ vs T for $\nu=0.5$. $+$: white noise, $\diamond$: accelerated, $*$: Boltzmann. The dotted line is the Carnahan and Starling relation $G_{CS}(\nu)$, given by~(\ref{CandS}), for $\nu=0.5$.} \label{GvsT} \end{figure} The substantial differences between $G_s(\nu)$, deduced from the equation of state, and $G_{CS}(\nu)$, predicted by the Carnahan and Starling relation (Eq.~(\ref{CandS})), comes from two sources. First, Eq.~(\ref{CandS}) may not give the correct value of the radial distribution function at the point of contact, $g(\sigma,\nu)$, for granular media. Second, $G_s(\nu)$ may not equal $\nu g(\sigma,\nu)$. Because Eq.~(\ref{Gg}) is the definition of $G$, this would amount to a change in the equation of state, which we assumed to be correct in the calculation of $G_s(\nu)$. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Bothgofr.ps}} \smallskip \caption[Radial distribution function for inelastic particles] {$g(r)$ at $\nu=0.5, T=1.05$ for (a) white noise ($r_k=1$) and (b) Boltzmann ($r_k=4$) forcing. The dot-dashed lines represent the value of $g$ given by the Carnahan and Starling relation Eq.~(\ref{CandS}) for $\nu=0.5$, while the dashed lines show $G_s(\nu)/\nu$. For white noise forcing, $g(\sigma,\nu)$ coincides with neither line, while for the Boltzmann bath, $g(\sigma,\nu)$ coincides with $G_s(\nu)/\nu$.} \label{gofrinelastic} \end{figure} To elucidate these two causes, we plot $g(r,\nu)$ for a run with white noise forcing and a run forced with a Boltzmann bath. Each plot also indicates the value of $g(\sigma,\nu)$ predicted by Eq.~(\ref{CandS}), as well as that predicted by Eq.~(\ref{Gg}), assuming $G(\nu)=G_s(\nu)$. For neither forcing type does~(\ref{CandS}), the Carnahan and Starling relation for $G_{CS}(\nu)$, properly predict $g(\sigma,\nu)$; rather, inelastic particles are more likely to be nearly in contact than elastic particles at the same density and temperature. Furthermore, while $G_s(\nu) = \nu g(\sigma,\nu)$ for the Boltzmann forcing, this does not hold for the white noise forcing or the accelerated forcing (not shown). Even though $g(\sigma,\nu)$ is larger than that predicted by the Carnahan and Starling relation, $G_s(\nu) < G_{CS}(\nu)$, indicating that the equation of state is incomplete. Recalling that the Boltzmann driving represents particles in contact with a highly randomizing bath, we conclude that the failure of the equation of state, (\ref{state}), is due to incomplete randomization of particle velocities through collisions, or in short, a breakdown of molecular chaos. \subsection{Velocity Correlations} Molecular chaos is the assumption that particle velocities are uncorrelated. Knowing the velocity of one of a pair of a colliding particle gives no information about the velocity of the other. In light of the behavior of $G$, and simulations of driven 1D and cooling 2D gases that showed strong velocity correlations~\cite{swift98}, we measure velocity-velocity correlation functions. Given two particles, labeled $1$ and $2$, ${\bf \hat{k}}$ is a unit vector pointing from the center of $1$ to the center of $2$. Particle $1$'s velocity then has a components $v_{1}^{||}$ parallel to and $v_{1}^{\perp}$ perpendicular to ${\bf \hat{k}}$; likewise for particle $2$. We define two correlation functions \begin{eqnarray} \langle v_{1}^{||} v_{2}^{||} \rangle &=& \sum v_{1}^{||} v_{2}^{||} / N_r,\\ \langle v_{1}^{\perp} v_{2}^{\perp} \rangle &=& \sum v_{1}^{\perp} v_{2}^{\perp} / N_r, \end{eqnarray} where the sums are over $N_r$ particles such that the distance between the two particles is within $\delta r$ of $r$. If particle velocities are uncorrelated, $\langle v_{1}^{||} v_{2}^{||} \rangle$ and $\langle v_{1}^{\perp} v_{1}^{\perp} \rangle$ will both give zero. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Vcorrelations2.ps}} \smallskip \caption[Velocity correlation functions]{Velocity correlations as a function of particle separation at $\nu = 0.5, T=1.05$ for: ($+$) white noise, ($\diamond$) accelerations, ($\times$) Boltzmann forcing, and ($\triangle$) elastic particles. Each curve is built from around 100 frames separated in time by 100 collisions per particle, and $\delta r = \sigma/10$. Both the elastic particles and the particles forced with the Boltzmann bath have essentially zero correlation over most of the range. The Boltzmann bath shows positive correlations only at very short range.} \label{vllvpp} \end{figure} The parallel and perpendicular velocity correlations are plotted in Fig.~\ref{vllvpp} for the three types of forcing and for elastic particles. Both for particles driven with white noise and accelerations, strong long-range velocity correlations are apparent, with more correlations produced by the accelerated forcing, consistent with its stronger deviations in the single particle velocity distribution and in $G$. These correlations are not small, reaching as much as $40\%$ of the temperature; typically, the perpendicular correlations are about one-half of the parallel correlations. Further, these correlations are long range --- they extend the full length of the system. The parallel correlations drop to zero at L/2, while the perpendicular correlations reach zero around $r=10\sigma$, and have a negative value but zero derivative at $L/2$. The long-range of nature of the correlation is not due to the size of the computational cell. Similar cell-filling correlations were observed in runs 4, 16, and 64 times larger~\cite{bizon99a}. For the Boltzmann forcing, some correlations are visible at very short range; inelastic collisions are trying to establish correlations, but before these correlations can be extended to long range they are wiped out by the thermal bath. \subsection{Loss Rate} The loss rate of temperature due to inelastic collisions, $\gamma$, divided by the rate calculated from kinetic theory, $\gamma_0$ (see Eq.~(\ref{gamma0eq})), is shown for the three forcing methods as a functions of $T$ in Fig.~\ref{gammaTpic}(a) and $\nu$ in Fig.~\ref{gammanupic}(a). For the calculation of $\gamma_0$, $G$ was taken from the equation of state measurements, and the average $e$ within a run was used. Most surprising is the increased loss rate over kinetic theory at relatively low temperature. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{gammaT.ps}} \smallskip \caption{Temperature dependence of loss rate with $\nu=0.5$. (a) $\gamma/\gamma_0$, where the kinetic theory result $\gamma_0$ [Eq.~(\protect{\ref{gamma0eq}})] assumes a velocity-independent restitution coefficient. (b) $\gamma/\gamma_e$, where $\gamma_e$ takes into account the velocity dependence of $e$. The symbols denote the type of forcing: $+$ White noise; $\diamond$ Accelerated; $\times$ Boltzmann. The dotted lines show ${{\sqrt{\pi}}\over{4\sqrt{T}}} \langle v_n^2\rangle_c / \langle v_n \rangle_c$ for the white noise and accelerated runs; see (\protect\ref{avggamma}).} \label{gammaTpic} \end{figure} The calculation of $\gamma$ requires only the evaluation of \begin{equation} \gamma = {{\sigma}\over{2}}\int\int {{1}\over{4}}(1-e^2) ({\bf v}_{12}\cdot \hat{\bf k})^3 f^{(2)}({\bf v}_1,{\bf v}_2) d\Omega d{\bf v}_1 d{\bf v}_2, \label{cintegral} \end{equation} where $d\Omega$ is the angle element. This expression simply averages the energy lost per collision, \begin{equation} {{1}\over{4}}(1-e^2) v_n^2 = {{1}\over{4}}(1-e^2) ({\bf v}_{12}\cdot \hat {\bf k})^2 \end{equation} over all possible collisions. The remaining factor of $({\bf v}_{12}\cdot \hat {\bf k})$ takes into account the fact that grains traveling towards one another more rapidly are more likely to collide during any given time interval. The kinetic theory result for the loss rate, Eq.~(\ref{gamma0eq}), follows from the collisional integral, Eq.~(\ref{cintegral}), under two assumptions: molecular chaos, which ought to be a more reasonable assumption at lower temperatures, and the independence of $e$ on the variables of integration, so that it is pulled out of the integral as a constant. At lower temperature, where the variation of $e$ with $v_n$ leads to a distribution of $e$ at a given temperature, the second assumption fails. As temperature drops still farther, the spread of $e$ reduces, since $e$ is bounded from above by $1$; at the very lowest $T$, $\gamma$ does approach $\gamma_0$. Substituting the function form of $e(v_n)$, Eq.~(\ref{restform}), into the collisional integral, Eq.~(\ref{cintegral}), assuming molecular chaos, and performing the integrations, we arrive at an equation for $\gamma = \gamma_e$ that takes into account the variation of $e$ with $v_n$: \begin{equation} \gamma_e={{4 \nu G \sqrt{T}}\over{\sigma^3 \pi^{3/2}}}\left[(1-e_0^2)(v_a^2+4T) \exp(-v_a^2/4T) + 4 I\right], \end{equation} where \begin{eqnarray} I &=& 2^{1+\beta} A T^{1+\beta/2} (\Gamma(2+\frac{1}{2}\beta)-\Gamma(2+\frac{1}{2}\beta,v_a^2/4T))\nonumber\\ &-& A^2 2^{2\beta} T^{1+\beta}(\Gamma(2+\beta)-\Gamma(2+\beta,v_a^2/4T)), \end{eqnarray} $\Gamma(a)$ is the gamma function, and $\Gamma(a,b)$ is the incomplete gamma function. In the limit that $v_a \rightarrow 0$, $\gamma_e \rightarrow \gamma_0$. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{gammanu.ps}} \smallskip \caption{(a) $\gamma/\gamma_0$ and (b) $\gamma/\gamma_e$ versus $\nu$ at $T=1.05$ The symbols denote the type of forcing: $+$ White noise; $\diamond$ Accelerated; $\times$ Boltzmann} \label{gammanupic} \end{figure} Figs.~\ref{gammaTpic}(b) and ~\ref{gammanupic}(b) show $\gamma/\gamma_e$ for the same simulations shown in Figs.~\ref{gammaTpic}(a) and ~\ref{gammanupic}(a). Taking the variations in $e$ into account removes the underestimation of $\gamma$. In the revised picture, $\gamma$ approaches $\gamma_e$ at low $T$ and at low $\nu$, but as either increases, $\gamma$ drops from the value predicted by $\gamma_e$. This is due to the velocity correlations produced by the inelasticity. Locally, particles are moving together, reducing collision velocities and collision frequencies, thereby reducing the loss rate; see Fig~\ref{PvsType}. For the Boltzmann forcing, velocity correlations are wiped out, and $\gamma$ is close to $\gamma_e$. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{PvcThermostat.ps}} \smallskip \caption[Probability distribution of relative velocities at collision]{Probability distribution of $v_c \equiv |{\bf v}_1 - {\bf v}_2|$, the magnitude of relative velocities at collision for different forcings ($+$) white noise, ($\diamond$) acceleration, ($\times$) Boltzmann, and ($\triangle$) for elastic particles, all at $\nu=0.5$ and $T=1.05$. The solid curve is the distribution predicted by uncorrelated collisions between particles chosen from Boltzmann distributions, $(1/2\sqrt{\pi T^3}) v_c^2 e^{-v_c^2/4T}$. Positive velocity correlation of nearby particles causes a reduction in the collision velocities, and hence a reduction in $\gamma$.} \label{PvsType} \end{figure} Writing $\gamma$ in terms of average quantities makes its dependence on the collision velocity more explicit. The loss rate is identically equal to the average energy lost per collision, $\langle \Delta E \rangle_c$, times the average collision frequency per volume $f/V$ \begin{equation} \gamma=\langle \Delta E \rangle_c f/V = {{1}\over{4}} (1-e^2)\langle v_n^2\rangle_c f/V, \label{name} \end{equation} where the c subscript denotes an average over collisions, and assuming again that $e$ is independent of collision velocity. Similarly, the virial equation of state, Eq.~(\ref{mdstate}), in terms of average quantities is \begin{equation} P=(4/\pi\sigma^2) \nu T(1+{{V\sigma}\over{d}}f{{1+e}\over{2}}\langle v_n \rangle_c). \label{avgG} \end{equation} Solving for $f$ in terms of $G$ from Eqs.~(\ref{state}) and ~(\ref{avgG}), and substituting this into Eq.~(\ref{name}) we obtain \begin{equation} \gamma={{(1-e^2)G}\over{\sigma}} {{\langle v_n^2\rangle_c}\over{\langle v_n\rangle_c}} nT. \label{avggamma} \end{equation} If the distribution of relative normal velocity at collision is equal to that predicted by molecular chaos, $P(v_n) = (1/2T) v_n \exp{-v_n^2/4T}$, then $\langle v_n^2\rangle_c = 4T$ and $\langle v_n\rangle_c = \sqrt{\pi T}$, so that $\gamma = \gamma_0$ is recovered. As seen in Fig.~\ref{PvsType}, however, the distribution of collision velocities is different from the molecular chaos values due to the presence of velocity correlations. To show that this deviation accounts for the remaining difference between $\gamma$ and $\gamma_e$, $\langle v_n^2\rangle_c$ and $\langle v_n\rangle_c$ were calculated in the simulations. Their ratio, normalized by the molecular chaos value $4 \sqrt{T/\pi}$ is plotted on Figs.~\ref{gammaTpic}(b) and ~\ref{gammanupic}(b) as dotted lines. Except where there are wide variations of $e$ within a given run (such as at high velocity in the accelerated runs), the change in the relative collision velocities tracks the change in $\gamma$. \section{Inhomogeneous Forcing and Transport Coefficients} \label{inhomogeneous} So far, we have been concerned with homogeneous forcing. By using applied forcing that varies spatially, we can induce inhomogeneous steady states. Then, by measuring fluxes, transport coefficients are calculated, and compared to kinetic theory. Inhomogeneous states have only been calculated for the accelerated forcings; measurements for the homogeneous state show that deviations from the elastic case are strongest in this case. \subsection{Thermal Conductivity} \label{thermalconductivity} Recall that with the accelerated forcing, the direction of the accelerations of particles fluctuated at a fixed rate, but the magnitude was always the same. To induce a thermal gradient in the simulation, we allow the magnitude of the acceleration to vary in space. Specifically, when the acceleration of a particle is to be rotated, the magnitude of its acceleration is given by \begin{equation} a_0(1-|(z-\frac{1}{2}L)/(\frac{1}{2}L)|), \label{gradforcing} \end{equation} {\it i.e.,} we apply a linear gradient in the forcing, dependent on one spatial direction ($z$), peaked in the center of the cell and falling to zero at the periodic boundary. In order to preserve the center of mass momentum, the partner particle receives the opposite acceleration, regardless of its position in the cell, as described in Section~\ref{accelerated}. Under this forcing, a stable thermal profile develops, as seen in Fig.~\ref{profilesK}. The system reaches a mechanical equilibrium; the pressure is nearly constant in $z$. Constant $P$ and varying $T$ imply varying $\nu$ through the equation of state, and this variation is observed. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Tandnuofz.ps}} \smallskip \caption[Temperature, density, and pressure profiles for inhomogeneous forcing]{ Profiles of $*$: T, $\Box$: $\nu$, and $\diamond$: $P/5$ for inhomogeneous forcing in which the magnitude of the acceleration depends linearly on the distance from $z/L = \frac{1}{2}$. The pressure is nearly constant, so that the variation in $T$ induces a variation in $\nu$ through the equation of state. For this run, the average solid fraction is $0.75$.} \label{profilesK} \end{figure} The cell is divided into 50 slabs along $z$ for measurement purposes; for each slab, snapshots allow calculation of the average $T$ and $\nu$ within. Because the pressure may in principle vary over space, its measurement using the virial fails. Pressure is measured at the interfaces between the slabs by keeping track of the normal ($z$) momentum flux through these boundaries, both due to particles freely traveling through them, and due to collisions between particles that lie in different slabs. In addition, the energy added due to the forcing and the energy lost due to inelastic collisions are separately accounted for each slab. The difference between the energy gain and loss in a given slab must, in a steady state, be made up for by the difference in energy flux through its two boundaries, so that the net rate of change of the energy in the slab is zero. Assuming that the energy flux through the line at $z=(0,L)$ is zero due to symmetry allows calculation of ${\bf q}_z$, the heat flux through each slab boundary; see Fig~\ref{flux}. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{qanddtdz.ps}} \smallskip \caption[Heat flux and temperature gradient]{The $\times:$ thermal gradient and $\triangle$: heat flux for the run shown in Fig.~\protect{\ref{profilesK}}.} \label{flux} \end{figure} Once ${\bf q}_z$ and $\partial T/\partial z$ are calculated, Fourier's heat law, Eq.~(\ref{fourierslaw}), can be used to calculate the thermal conductivity $\kappa$. The results of many simulations, holding the average $\nu$ at $0.75$, but varying $a_0$ in Eq.~(\ref{gradforcing}) and therefore the size of the thermal gradient, are shown in Fig.~\ref{Conductivity}, are compared to the result of Enskog theory, as given by Eq.~(\ref{conductivityenskog}). At low temperatures, Enskog theory does a good job predicting $\kappa$. However, as the temperature increases and $e(v_n)$ decreases, Enskog theory does worse; at the highest temperatures, Enskog theory overestimates the thermal conductivity by a factor of 2. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Kappa.ps}} \smallskip \caption[Deviation from Enskog theory for thermal conductivity]{Ratio of $\kappa$ measured from simulations to $\kappa_0$ from Enskog theory (Eq.~(\ref{conductivityenskog})). Each symbol denotes a different run, but for each run, the average solid fraction is $0.75$.} \label{Conductivity} \end{figure} Note that this calculation does not test Fourier's Law; rather we assume that Fourier's law is correct, and use it to calculate $\kappa$. Analysis based on closures of the Boltzmann equation predict a term in the heat flux proportional to the density gradient~\cite{brey97a}. If such a term had a sizeable magnitude and were ignored, it would cause a reduction in the observed $\kappa$. \subsection{Shear Viscosity} Spatial inhomogeneity in the magnitude of the forcing led to a stationary inhomogeneous temperature field, allowing measurement of heat flux and thermal conductivity; spatial inhomogeneity in the mean of the forcing leads to a stationary inhomogeneous velocity field, allowing measurement of the momentum flux and the shear viscosity. In particular, particle accelerations are chosen according to \begin{eqnarray} a_y &=& a_0(0.01 \sin{(2\pi z/L)} + \psi_0)\\ a_z &=& a_0 \psi_1 \end{eqnarray} where $\psi_0$ and $\psi_1$ are numbers chosen randomly from a Gaussian distribution with zero mean and standard deviation of 1. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{SinWave.ps}} \smallskip \caption[Velocity profile]{Average velocity in the $y$ direction as a function of $z$ at $T=0.21$, $\nu=0.6$. The solid line is sinusoidal.} \label{velocity} \end{figure} This forcing produces steady states with velocity, temperature, density, and stress fields like that shown in Fig~\ref{velocity}. The velocity profile is nearly sinusoidal, and the temperature, pressure, and solid fraction are essentially independent of $z$. In the simulations discussed so far, we have only considered the scalar quantity $T = \langle (v - \langle v \rangle )^2\rangle / D$, where $D$ is the number of dimensions and the $\langle\rangle$ denote averages over particles. This temperature is more generally the trace of the temperature tensor: \begin{equation} T_{ij}=\langle (v_{i} - \langle v_i \rangle )(v_{j}-\langle v_j \rangle)\rangle, \end{equation} where $i,j$ range over the directions, and $v_i$ denotes the $i$-th component of the velocity. In principle, $T_{yy}$ need not equal $T_{zz}$ if the rate at which fluctuational energy is traded between the directions is slower than the rate at which it is added anisotropically; such is the case in a vertically oscillated granular layer, where vertical fluctuational energy can be twice that of the horizontal. Introducing a bias in the acceleration of only 1\%, however, does not introduce anisotropy into $T$; for larger shear rates this is not the case. To study the simplest case, we restrict our simulations to biases of 1\%. In section~\ref{thermalconductivity} we described calculation of the pressure by measuring the normal momentum flux through planes. Measuring the tangential flux through the planes, and introducing a second set of planes orthogonal to the first, allows calculation of the full 4-component pressure tensor. As seen in Fig.~\ref{stressdifference}, the pressure tensor is anisotropic even though $T$ is isotropic; the anisotropy increases as $T$ increases, or $e$ decreases. For $e\approx 1$, the stress difference is approximately proportional to $1-e$, but the variation in $e$ within a given run probably plays a role, as it did in the loss rate; see Fig.~\ref{stressdifference}. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{StressDifference.ps}} \smallskip \caption[Stress difference vs. e]{Normal Stress difference divided by pressure as a function of $1-e$. The line has slope 1.} \label{stressdifference} \end{figure} For each run at fixed $T$, we can test Newton's viscosity law, \begin{equation} P_{yz} = - \mu {{\partial v_y}\over{\partial z}} \label{viscouslaw} \end{equation} where the viscosity $\mu$ is a constant of proportionality. A typical result is shown in Fig.~\ref{newton}, where the linear relation of Eq.~(\ref{viscouslaw}) is shown to hold. The slope of these curves then provide values for $\mu$ which can be compared to the Enskog result from Eq.~(\ref{viscosityenskog}); the results are shown in Fig.~\ref{Viscosity}. \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Newton.ps}} \smallskip \caption[Stress versus strain rate]{The $yz$ component of the pressure tensor versus the corresponding velocity derivative for the run of Fig.~\ref{velocity}. The slope of the best fit line is $-\mu$.} \label{newton} \end{figure} \begin{figure} \epsfxsize=.9\columnwidth \centerline{\epsffile{Viscosity.ps}} \smallskip \caption[Viscosity compared to the Enskog values]{Viscosity, normalized by the Enskog value $\mu_0$, as a function of T, for $\nu=0.6$.} \label{Viscosity} \end{figure} Unlike the loss rate and thermal conductivity, we find that the Enskog theory underestimates the shear viscosity at lower temperatures. However, the trend of decreasing transport with increasing T is the same. Even for elastic particles, Enskog theory is not expected to work to arbitrarily high solid fractions; as density increases, deviations from Enskog theory are expected. As the inelasticity of particles increases, velocity correlations increase, reducing collisional momentum transport and lowering the viscosity. \section{Conclusion} \label{conclusion} Volumetric driving of granular media leads to nearly stationary states that are amenable to comparison with kinetic theory, allowing us to test the six points of kinetic theory listed in Sec.~\ref{intro}. Given that many of our simulations involve coefficients of restitution that are not near 1, the general level of agreement with kinetic theory is surprisingly good, suggesting that continuum approaches to dissipative granular media are capable not only of qualitative, but also quantitative descriptions of real systems. We now discuss each of the six points in turn. \emph{Single particle distribution functions are nearly Boltzmann}. For all forcing types, temperatures (coefficients of restitution), and densities, the single particle distribution functions are close to Boltzmann distributions (Fig.~\ref{oneparticledist}). The deviations from Gaussian (Fig.~\ref{ratio}) are consistent with but much smaller than those seen in experiments on thin ($< 1$ layer deep) oscillated granular media~\cite{olafsen98}. In those experiments, deviations appear to be due to spontaneous spatial variations in temperature that are not taken into account in the analysis~\cite{urbachpersonal}; such variations become large for cooling, unforced granular media. The smaller deviations exhibited in our simulations may represent smaller temperature fluctuations. \emph{Particle velocities are correlated.} Standard kinetic theory assumes molecular chaos: particle velocities are uncorrelated. In our simulations, as in simulations of cooling granular media~\cite{orza98}, strong velocity correlations (Fig.~\ref{vllvpp}) with a characteristic vortex structure~\cite{bizon99a} develop. In the steady state, these correlations extend over the entire cell. Molecular chaos is not required for kinetic theory; a closure for the kinetic equation is critical. Similar considerations have lead van Noije and Ernst to apply ring kinetic theory~\cite{ernst71}, which allows for correlations in particle velocities, to a granular system~\cite{vannoije98b}. \emph{Spatial correlations are stronger than predicted by the Carnahan and Starling relation (Eq.~(\ref{CandS}))}. Even for simulations using the Boltzmann bath, in which velocity correlations are removed, the effect of inelasticity is to increase the amount of spatial correlation at $r=\sigma$ (Fig.~\ref{gofrinelastic}(b)). In other words, particles are more likely in the inelastic case to be close to one another than elastic particles in the same thermodynamic state. The size of this effect is dependent on the inelasticity, but can be greater than $15\%$. \emph{The equation of state overestimates the collisional contribution to pressure because it ignores velocity correlations.} The factor in the equation of state describing the contribution from collisions, $G_s(\nu)$, as calculated from the measurement of pressure and the equation of state, is smaller for white noise and acceleration forcings than that predicted by the Carnahan and Starling relation (Eqs.~(\ref{CandS}) and ~(\ref{Gg})), denoted $G_{CS}(\nu)$ (Fig.~\ref{Ginelastic}). In turn, $G_{CS}(\nu)$ is smaller than the actual $G(\nu) \equiv \nu g(\nu)$, as discussed in the previous paragraph. Because velocity correlations were ignored in the derivation of the equation of state (\ref{state}), the pressure due to collisions that $G$ describes is overestimated. To some degree, the increased positional correlation and increased velocity correlation work against one another; the first increases the collision frequency, while the latter decreases it. The net result is that the $G_s(\nu)$ from the pressure measurement is closer to Carnahan and Starling (\ref{CandS}) $G_{CS}(\nu)$ than if there were only velocity correlations (Fig.~\ref{gofrinelastic}(a)). \emph{Newton's stress law works well for low stress.} Even at the highest inelasticity, $e=0.7$, no deviations from a linear relation between stress and strain rate (Eq.~(\ref{newtonslaw})) were observed (Fig.~\ref{newton}). However, in order to keep the temperature isotropic, we have limited ourselves to cases in which $v^2 < T$; many flows of interest are supersonic, with average velocities much larger than $\sqrt{T}$. Because $\kappa$ depends on $T$, and therefore on position, Fourier's heat law was not tested in the same manner that Newton's viscosity law was. \emph{Temperature loss rate, $\gamma$, and thermal conductivity, $\kappa$, are reduced by inelasticity, while shear viscosity, $\mu$, is predicted relatively well by Enskog theory.} For increasing inelasticity or density in homogeneously forced runs, velocity correlations also increase. As a consequence, the relative collision velocity decreases (Fig.~\ref{PvsType}), leading to a reduction in the temperature loss rate due to inelastic collisions (Figs.~\ref{gammaTpic}(b) and ~\ref{gammanupic}(b)). In the most severe cases examined, once corrected for variations in $e$, this deviation could be as high as $20\%$. Inhomogeneously forced runs allowed calculation of $\mu$, and, assuming Fourier's law, $\kappa$; $\mu$ never deviated from the prediction of Enskog theory by more than $15\%$ (Fig.~\ref{Viscosity}), while $\kappa$ was found to be smaller than predicted by a factor of two for high inelasticities (Fig.~\ref{Conductivity}). This differential success suggests that velocity correlations, which should be present in both cases, are not responsible for the large reduction in $\kappa$. Rather, the inelasticity itself may be the cause. Enskog theory, applied to granular media, assumes that $e \approx 1$; the values of $\kappa_0$ and $\mu_0$ are the same as those for elastic particles. When grains collide, energy is dissipated, so that the amount transported collisionally is necessarily reduced as $e$ decreases. On the other hand, momentum is still conserved, so that $\mu$ is relatively unaffected. Also,some deviation from Enskog theory is possible due to a term in the heat flux proportional to density gradients. The results described above suggest a number of avenues for future research. First, measurements of viscosity should be extended into the supersonic regime. Second, more extensive calculations of thermal conductivity at different densities, and with different spatial forcings, should be undertaken to ascertain the role of density gradients. Third, time-dependent calculations should be performed, allowing measurement of the bulk viscosity. Such calculations should also provide measurements of the frequency dependence of the transport coefficients, which may be relevant for oscillated granular media. Fourth, the granular continuum equations can be used to perform stability calculations on problems such as vertically oscillated granular media~\cite{mythesis,bizon99b}. Finally, new forcing geometries should be explored, allowing direct comparison between particle simulations, continuum theories, and experiments. \section{Acknowledgments} We deeply thank Professor Jim Jenkins for helping us penetrate granular kinetic theory. This work was supported by the Department of Energy Office of Basic Energy Sciences.
\section{Introduction \protect\\} \label{sec:intro} The possibility of the existence of multiquark states was predicted by QCD inspired models \cite{jaf,muld}. These works initiated a lot of experimental searches for six-quark states (dibaryons). Usually one looked for dibaryons in the NN channel. In the present work we will consider supernarrow dibaryons (SNDs), the decay of which into two nucleons is forbidden by the Pauli exclusion principle \cite{fil1,akh,fil2,ger,alek,alek1}. In the NN channel such dibaryon states correspond to even singlets and odd triplets at the isotopic spin $T=0$, and odd singlets and even triplets at $T=1$. These dibaryons with the mass \mbox{$M < 2m_{N}+m_{\pi}$} ($m_N (m_{\pi}$) is the nucleon (pion) mass) can decay into two nucleons, mainly emitting a photon. This is a new class of dibaryons with decay widths \mbox{$\le 1$keV}. The contribution of such dibaryons to strong interaction processes of hadrons is small. However, their contribution to electromagnetic processes on light nuclei may exceed the cross section for the process under study out of range of the dibaryon resonance by several orders of magnitude \cite{fil2,alek,alek1}. The experimental discovery of such states would have important consequences for particle and nuclear physics. In the frame of the MIT bag model, Mulders et al. \cite{muld} calculated the masses of different dibaryons, in particular, NN-decoupled dibaryons. They predicted dibaryons $D(T=0;J^P =0^{-},1^{-},2^{-};M=2.11$ GeV) \ and $D(1;1^{-};2.2$ GeV) corresponding to the forbidden NN states $^{13}P_J$ and $^{31}P_1$. However, the dibaryon masses obtained exceed the pion production threshold. Therefore, these dibaryons can decay into the $\pi NN$ channel. The possibility of the existence of dibaryons with masses $M<2m_N+m_\pi $ was discussed by Kondratyuk et al. \cite{kon1} in the model of stretched rotating bags, taking account of spin-orbital quark interactions. In the frame of the chiral soliton model, Kopeliovich \cite{kop} predicted that the masses of $D(T=1,J^P=1^+)$ and $D(0,2^+)$ dibaryons exceeded the two-nucleon mass by 60 and 90 MeV, respectively. These values are lower than the pion production threshold. Unfortunately, all results obtained for the dibaryon masses are model dependent. Therefore, only an experiment could answer the question about the existence of SNDs and their masses. In the work \cite{bilg} the existence of a dibaryon, called $d'$, with quantum numbers $T=even$ and $J^P=0^-$ which forbid its decay into two nucleons, and with the mass $M=2.06$ GeV, and the decay width $\Gamma_{\pi NN}=0.5$ MeV, has been postulated to explain the observed resonance-like behavior in the energy dependence of the pionic double charge exchange on nuclei at an energy below the $\Delta$-resonance. However, there is a more convential interpretation of these data. It was shown \cite{nus} in the frame of the distorted-wave impulse approximation that such a peak arises nuturally because of the pion propogation in the sequential process, in which pion double charge exhange occurs through two successive $\pi N$ charge exchange reactions on two neuterons. In the work \cite{khr} dibaryons with exotic quantum numbers were searched for in the process $pp\to pp \gamma\g$. The experiment was performed with a proton beam from the JINR phasotron at the energy of 198 MeV. Two photons were detected in the energy range between of 10 and 100 MeV. Some structure was observed in the photon energy spectrum, which was attributed to the exotic dibaryon with the mass $M\approx 1920$ MeV. The statistical significance of the effect is about 8$\sigma$. However, a big width (FWHM=$31.3\pm 5.0$ MeV) of the observed structure was obtained, and an additional careful experimental study of this reaction is needed in order to understand the nature of this structure. It is necessary to measure the photon energy spectrum with a higher energy resolution and a higher statistical precision. Moreover, measurements at two different incident proton energies are needed to check whether the structure observed is a SND. On the other hand, an analysis \cite{cal} of the Uppsala proton-proton bremsstrahlung data looking for the presence of a dibaryon in the mass range from 1900 to 1960 MeV only gave the upper limits of 10 and 3 nb for the dibaryon production cross section at proton beam energies of 200 and 310 MeV, respectively. This result disagrees with the data \cite{khr}. Earlier we have carried out measurements of missing mass spectra in the reaction $pd\to pX$ \cite{konob} using a double-arm spectrometer at the Moscow Meson Factory. The narrow peak in the missing mass spectrum at 1905 MeV with a width equal to the experimental resolution of 7 MeV has been observed in this experiment. As will be shown below, the nucleons and the deuteron from the decay of the dibaryons under consideration into $\gamma NN$ and $\gamma d$ have to be emitted in a narrow angular cone with respect to the direction of motion of the dibaryon. The size of this cone depends also on the quantum numbers of the dibaryon. So a detection of the scattered proton in coincidence with the proton (or the deuteron) from dibaryon decay at correlated angles gives a good possibility to separate the SNDs from the background and to determine their quantum numbers. In the present paper we give a more detailed study of the reactions $pd\to p+pX_1$ and $pd\to p+dX_2$ (where $X_1$ and $X_2$ are undetected particles in this experiment) with the aim to search for SNDs at the Moscow Meson Factory using an improved facility. We will consider the following dibaryons: $D(T=0,J^P=0^+)$, $D(0,0^-)$, $D(1,1^+)$, and $D(1,1^-)$. It is worth noting that the state $(T=1, J^P=1^-)$ corresponds to the states $^{31}P_1$ and $^{33}P_1$ in the NN channel. The former is forbidden and the latter is allowed for a two-nucleon state. In our work we will study the dibaryon $D(1,1^-)$, a decay of which into two nucleons is forbidden by the Pauli principle (i.e. $^{31}P_1$ state). The contents of the paper are the following. In Sec.\ \ref{sec:2} the decay widths of SNDs are given and the cross sections of these dibaryon production in the processes $pd\to pD$ are calculated. A description of the experimental setup is in Sec.\ \ref{sec:setup}. In Sec.\ \ref{sec:result} the results of the measurements are presented. The results of a Monte Carlo calculation and an analysis of the experimental data obtained are presented in Sec.\ \ref{sec:analys}. The main conclusions are given in Sec.\ \ref{sec:concl}. \section{Cross sections of the supernarrow dibaryon production in the reaction $\lowercase{pd}\to \lowercase{p}D$ \protect\\} \label{sec:2} In the process $pd\to pD$, SNDs can be produced only if the nucleons in the deuteron overlap sufficiently, such that a six-quark state with deuteron quantum numbers can be formed. In this case, an interaction of a photon or a meson with this state can change its quantum numbers so that a metastable state is formed. Therefore, the probability of the production of such dibaryons is proportional to the probability $\eta$ of the six-quark state existing in the deuteron. The magnitude of $\eta$ can be estimated from the deuteron form factor at large $Q^2$ (see, for example, \cite{bur}). However, the values obtained depend strongly on the model of the form factor of the six-quark state over a broad region of $Q^2$. Another way to estimate this parameter is to use the discrepancy between the theoretical and experimental values of the deuteron magnetic moment \cite{kim,kon2}. This method is free from the restrictions quoted above and gives $\eta\le 0.03$ \cite{kon2}. We assume here that $\eta$ and the probability of the full overlap of nucleons in the virtual singlet state $^{31}S_0$ are equal to 0.01. Since the energy of nucleons, produced in the decay of the dibaryons under study with $M<2m_N +m_{\pi}$, is small, it may be expected that the main contribution to a two-nucleon system should come from $^{13}S_1$ (deuteron) and $^{31}S_0$ (virtual singlet) states. The results of calculations of the decay widths of the dibaryons into $\gamma d$ and $\gamma NN$ on the basis of such assumptions are listed in Table \ref{table1}. As shown in \cite{alek1} the main contribution to the dibaryon photoproduction from the deuteron is given by the one-pion exchange diagram. Therefore, we will describe the production of SNDs in the process $pd\to p D$ using the one meson exchange diagram also. The production of the isoscalar dibaryons $D(0,0^+)$ and $D(0,0^-)$ will be calculated with the help of the pole diagram with $\omega $-meson exchange between the incident proton and the deuteron. The vertices $D(0,0^+)\to\omega d$ and $D(0,0^-)\to\omega d$ can be written as \begin{equation} \Gamma_{D(0,0^+)\to\omega d}=\frac{f_1}{M}\sqrt{\eta} \epsilon_{\mu\nu\lambda\sigma}G^{\mu\nu}\Phi^{\lambda\sigma}_1, \end{equation} \begin{equation} \Gamma_{D(0,0^-)\to\omega d}=\frac{f_2}{M}\sqrt{\eta} G_{\mu\nu}\Phi^{\mu\nu}_1, \end{equation} where $G_{\mu\nu}=v_{\mu}r_{\nu}-r_{\mu}v_{\nu}$, $\Phi_{1\mu\nu}=w_{\mu}p_{\nu}-p_{\mu}w_{\nu}$. Here $v$ and $w$ are the four-vectors of the deuteron and $\omega$ meson polarization, $r$ and $p$ are their four-momenta. As a result of the calculations we have \begin{equation} \frac{d\sigma_{pd\to pD(0^+)}}{d\Omega }= \frac{f_1^2}{4\pi }F_0R_{(+)}, \end{equation} \begin{equation} \frac{d\sigma_{pd\to pD(0^-)}}{d\Omega }= \frac 14\frac{f_2^2}{4\pi }F_0R_{(-)}, \end{equation} where \[ F_0=\frac{g_{\omega NN}^2}{4\pi }\frac{F}{(t-m_{\omega}^2)^2}, \] \[ F=\frac 83\eta \frac{p_2^2}{M^2}\frac{1} {m_dp_1\left[\left( m_d+E_1\right) p_2-p_1E_2\cos \theta \right]}, \] \begin{eqnarray*} R_{(-)}&=&-\frac14 \{2t[(s-m_d^2-m^2_N)(s+t-M^2-m^2_N)+tm_d^2] \\ &&+(M^2-m_d^2-t)^2(2m^2_N+t)\}, \end{eqnarray*} \[ R_{(+)}=R_{(-)}+tm_d^2(2m^2_N+t), \] \[ t=2(m^2_N-E_1E_2+p_1p_2\cos\theta), \] $E_1,p_1(E_2,p_2)$ are the energy and momentum of the initial (final) proton, $\theta $ is the angle of the scattered proton emission, $m_d$ is the deuteron mass, $g_{\omega NN}^2$ is the constant of the $\omega N$ interaction, $f_{1,2}^2/4\pi$ are the coupling constants in the vertices $D(0,1^{+})+\omega \to D(0,0^+)$ and $D(0,1^+)+\omega \to D(0,0^-)$, respectively, and $D(0,1^+)$ is the six-quark state with quantum numbers of the deuteron. The isovector dibaryons $D(1,1^+)$ and $D(1,1^-)$ can be formed mainly as a result of pion exchange between the incident proton and the deuteron. The vertices $D(1,1^+)\to \pi+d$ and $D(1,1^-)\to \pi+d$ are written as \begin{equation} \Gamma_{D(1,1+)\to\pi d}=\frac{g_1}{M}\sqrt{\eta}\epsilon_{\mu\nu\lambda\sigma} G^{\mu\nu}\Phi_2^{\lambda\sigma}, \end{equation} \begin{equation} \Gamma_{D(1,1-)\to\pi d}=\frac{g_2}{M}\sqrt{\eta} G_{\mu\nu}\Phi_3^{\mu\nu}, \end{equation} where $\Phi_{2\mu\nu}=\varepsilon_{1\mu}q_{1\nu}-q_{1\mu}\varepsilon_{1\nu}$, $\Phi_{3\mu\nu}=\varepsilon_{2\mu}q_{2\nu}-q_{2\mu}\varepsilon_{2\nu}$, and $\varepsilon_{1(2)}$ and $q_{1(2)}$ are the four-vectors of polarization and four-momenta of the $D(1,1^+)$ ($D(1,1^-)$), respectively. The calculations yield the following expressions for the cross section of these dibaryon production \begin{equation} \frac{d\sigma_{pd\to pD(1^+)}}{d\Omega }=\frac{g_1^2}{4\pi }F_1 [(M^2+m_d^2-t)^2-4m_d^2M^2], \label{one} \end{equation} \begin{equation} \frac{d\sigma_{pd\to pD(1^-)}}{d\Omega}=\frac14\frac{g_2^2}{4\pi}F_1 [(M^2+m_d^2-t)^2+2m_d^2M^2], \label{two} \end{equation} where \[ F_1=-\frac14\frac{g_{\pi NN}^2}{4\pi}\frac{tF}{(t-m_{\pi}^2)^2}, \] and $g_{1,2}^2/(4\pi )$ are the coupling constants in the vertex for the transition of the six-quark state $D(0,1^+)$ into the dibaryons under consideration via the pion exchange. For a numerical calculation of the dibaryon contribution let us assume that $g_{\pi NN}^2/4\pi =14.6$, $g_{\omega NN}^2/4\pi =19.2$, $\eta =0.01$. The constants $g_{1,2}^2/4\pi $ and $f_{1,2}^2/4\pi $ are unknown. They are strong interaction constants. In order not to overestimate the contributions of the dibaryons we put these constants equal to 1. The results of the calculation of the SND production cross sections in the process $pd\to pD$ at the kinetic energy of the incident proton $T_1=E_1-m=305$ MeV and the emission angle of the scattered proton $\theta =70^{\circ}$ as a function of the dibaryon mass $M$ are shown in Fig.~\ref{fig1}. Here the solid , dashed, dotted and dash-dotted lines correspond to $D(0,0^+)$, $D(0,0^-)$, $D(1,1^+)$, and $D(1,1^-)$ dibaryon production, respectively. As indicated above, the SNDs decay mainly emitting a photon. Therefore, if we limit ourselves to the investigation of $pd$ interactions with the photon in the final state, then the contribution of such dibaryons will essentially exceed the cross sections of the background processes. On the other hand, the special choice of the $pd\to pX$ kinematics allows us to allocate the area where the contribution of the SNDs dominates, even without detection of the final photon. \section{Experimental setup \protect\\} \label{sec:setup} The reaction $p+d\to p+X$ was studied at the proton accelerator of the Moscow Meson Factory at 305 MeV. A proton beam with an average effective intensity $\sim 0.1$ nA bombarded alternatively CD$_2$ and $^{12}$C targets of 0.14 and 0.18 g/cm$^2$, respectively. The $pd$ reaction contribution was determined by subtracting the $^{12}$C spectrum from the CD$_2$ spectrum. The exposition time of the experiment using CD$_2$ as a target was 100 hours. This period consisted of two runs corresponding to two different kinematic conditions in order to avoid a systematic error. The charged particles produced in the reaction in question were detected at different correlated angles by the Two Arms Mass Spectrometer (TAMS). Our layout is shown schematically in Fig.~\ref{fig2}. The left movable spectrometer arm, which is a single telescope $\Delta E-\Delta E-E$, was used to measure the energy and the time of flight of the scattered proton at a fixed emission angle $\theta_L\equiv\theta$. A remote control drive allowed to place this arm at various angles between $65^{\circ}$ and $85^{\circ}$. In the present experiment TAMS detected the scattered proton at the angle $\theta_L= 72.5^{\circ}$ (and $70^{\circ}$ in another run) in coincidence with the second charged particle (either $p$ or $d$) from the decay of the particle $X$ during 55 (and 45) hours. The detection of the second charged particle in the right arm at angles close to the emission angle of particle $X$ with mass $M$ allows to suppress essentially the contribution of background processes and increase the relative contribution of a possible SND production. The right arm consisted of three telescopes which were located at $\theta_R=33^{\circ}$, $35^{\circ}$ and $37^{\circ}$. These angles correspond to directions of the emission of the dibaryons with a certain mass. Each telescope included two thin plastic scintillation $\Delta E$ detectors ($4\times 4\times 0.5$ cm$^3$) and one $E$ detector of thick plastic ($5\times 5\times 20$ cm$^3$, used in the right arm) or BGO crystal ($5\times 5\times 8$ cm$^3$, used in the left arm). Each detector was viewed by a PMT-143 phototube coupled to specially developed fast electronic modules. A trigger was generated by 4-fold coincidence of the two $\Delta E$ detector signals of the left arm combined with those of any telescope of the right arm. A time resolution better than 0.5 ns was achieved, which allowed to suppress essentially an accidental coincidence background. The scattered proton $E$ signals, selected in the coincidence, formed the energy spectrum and accordingly the missing mass spectrum. Each useful event including two time of flights and two energies were stored event by event and then analyzed off line. The elastic $pd$ scattering was measured at various angles of the spectrometer arms. The proton energy spectra obtained were used to calibrate the spectrometer in the proton energy and, accordingly, in the missing mass. The missing mass resolution of the spectrometer was $\Delta M\approx 3$ MeV, the angular resolution was $1^{\circ}$. \section{Results of the measurements \protect\\} \label{sec:result} The experimental missing mass spectra obtained with the deuteried polyethylene target are shown in Figs.~\ref{fig3}(a,b,c). Each spectrum corresponds to a certain combination of outgoing angles of the scattered proton and the second charged particle. These combinations in Figs.~\ref{fig3}(b,c) are consistent with the shift of the emission angle $\theta_R$ of the dibaryon with the given mass when the angle $\theta_L$ changes from $70^{\circ}$ to $72,5^{\circ}$. As is evident from Figs.~\ref{fig3}(a) and \ref{fig3}(b), a resonance-like behavior of the spectra is observed for the CD$_2$ target in two mass regions at $1905\pm 2$ MeV and $1924\pm 2$ MeV. On the other hand, the experiment with the carbon target resulted in a rather smooth spectra \cite{izv}. This smoothness is caused by an essential increase of the contribution of background reactions in the interaction of the proton with the carbon. It is worth noting that the resonance structure at 1905 MeV is observed in both runs at $\theta_L=70^{\circ}$ and $\theta_L=72.5^{\circ}$ and at different setup modifications. This essentially decreases the possibility of a random origin of the observed structure. The results obtained at different $\theta_L$ were summed up to improve statistics. The mass spectra obtained in Figs.~\ref{fig3}(a,b,c) were interpolated by second order polynomials (for the background) plus Gaussians (for the peaks in Figs.~\ref{fig3}(a) and \ref{fig3}(b)). Let us regard Fig.~\ref{fig3}(b). In this case the interpolation gave $\chi^2/(n-5)=0.81$. We determine the number of standard deviations (SD) as \[ \frac{N_{eff}}{\sqrt{N_{eff}+N_{b}}}, \] where $N_{eff}$ is the number of events above the background curve and $N_{b}$ is the number of events below this curve. Taking 5 points for the peak at $M=1905$ MeV, we have \mbox{$65/\sqrt{185}=4.8$ SD}. That corresponds to the probability of statistic fluctuations $P$ equal to $2.7\times 10^{-5}$ \cite{stat}. Using the same procedure for the peak at $M=1924$ MeV in Fig.~\ref{fig3}(a) results in 4.9 SD and $P=1.6\times 10^{-5}$. If, for the number of SD, we use the expression \cite{tat} \[ \frac{\sum_{i}^{n=5} (N_{ti}-N_{bi})/\sigma_i^2} {\sqrt{\sum_i^{n=5} 1/\sigma_i^2}} \] where $N_{ti}$, $N_{bi}$, and $\sigma_i$ correspond, respectively, to total, background, and error bar data, then we obtain 4.2 S.D. for the peak at $M=1905$ MeV and 4.4 S.D. for the peak at $M=1924$ MeV. The widths of both observed peaks correspond to the experimental resolution (3 MeV). The peak at 1924 MeV was obtained only for one spectrum close to the upper limit of the missing mass. In the other cases this mass position was beyond the range of measurement. Therefore, in the present work we analyze in more detail the peak at 1905 MeV only. The experimental missing mass spectra in the range of 1872--1914 MeV, after subtracting the carbon contributions, are shown in Figs.~\ref{fig4}(a,b,c). As seen from Figs.~\ref{fig3} and \ref{fig4}, the resonance behaviour of the cross section exhibits itself in a limited angular region. \section{Analysis of the experimental data \protect\\} \label{sec:analys} If the observed structure at $M=1905$ MeV corresponds to a dibaryon decaying mainly into two nucleons, then the expected angular cone size of emitted nucleons would be about 50$^{\circ}$. Moreover, the angular distributions of the emitted nucleons are expected to be very smooth in the angular region under consideration. Thus, even assuming that the dibaryon production cross section is equal to that of elastic scattering (40 $\mu b$/sr) its contribution to the missing mass spectra in Figs.~\ref{fig3}(a,b,c) would be nearly the same and would not exceed 1--2 events. Hence, the peaks found are hardly interpreted as a manifestation of the formation and decay of such states. It was shown in \cite{fil2} that the decay of the SND into $\gamma NN$ had to be characterized in the rest frame by a narrow peak near the maximum photon energy in the probability distribution of the dibaryon decay over an emitted photon energy. It leads to an essential limitation of the outgoing nucleon angles. On the other hand, if such a dibaryon decays into $\gamma d$, the emitted deuteron angles are limited by the following condition: $\sin\theta_d\le M p^*_d/(m_d p_D)$, where $p_D$ is the momentum (in the lab. syst.) of the dibaryon and $p^*_d$ is the momentum (in the c.m.s.) of the deuteron. Using the Monte Carlo simulation we estimated the contribution of the SNDs with different quantum numbers and $M$=1905 MeV to the mass spectra at various angles of the left and right arms of our setup. The production cross section and branching ratio of these states were taken according to the calculations presented in Sec.\ \ref{sec:2} and the proton beam current was assumed to be equal to 0.1 nA. The results obtained for different quantum numbers and decay modes of the dibaryons are listed in Table \ref{table2}. This calculation showed that the angular cone of protons and deuterons emitted from a certain dibaryonic state can be rather narrow. The axis of this cone is lined up with the direction of the dibaryon emission. Therefore, placing the right spectrometer arm at the expected angle of the dibaryon emission we increase essentially the signal-to-background ratio. Figs.~\ref{fig5}(a,b) exhibit the angular distributions of charged particles (either $p$ or $d$) from the decay of the dibaryon $D(1,1^+)$ (Fig.~\ref{fig5}(a)) and $D(0,0^+)$ (Fig.~\ref{fig5}(b)). The solid and dashed curves correspond to the emission angles of the scattered proton equal to $70^{\circ}$ and $72.5^{\circ}$, respectively. The solid vertical lines show the location of the right arm detectors. These figures demonstrate the dependence of the angular distribution of the charged particles under consideration on the quantum numbers of the dibaryons. In Figs.~\ref{fig4}(a,b,c), the experimental spectra are compared with the predicted yields normalized to the maximum of the measured signal in Fig.~\ref{fig4}(b). The solid and dashed curves in these figures correspond to the states with the isospin $T$=1 and $T$=0, respectively. The background is described by a second order polynomial. As is seen from these figures and Table \ref{table2}, the ratios of the calculated yields to the given spectra are expected to be $0.3:1: 0.7$, if the state at 1905 MeV is interpreted as an isovector dibaryon ($D(T=1,J^P=1^{+})$ or $D(1,1^-)$). This is in agreement with our experimental data within the errors. On the other hand, the signals from isoscalar dibaryons ($D(0,0^+)$ or $D(0,0^-)$) could be observed in Figs.~\ref{fig4}(b,c) with the same probability. So, the dibaryon with $M=1905$ MeV is most likely to have the isospin equal to 1. In order to estimate the dibaryon production cross section, we use normalization to the elastic $pd$ scattering and then, comparing the theoretical predictions for the yields of one of the isovector dibaryons with the experimental data, find that the differential cross section of this dibaryon production is equal to $8\pm 4\;\mu b/sr$. That allows us to evaluate the product of unknown constants in Eqs.\ (\ref{one}) and (\ref{two}). If it is a dibaryon $D(1,1^+)$, then $\eta g_1^2/4\pi=(0.44\pm 0.22)\times 10^{-2}$. But if it is a dibaryon $D(1,1^-)$, we have $\eta g_2^2/4\pi=(0.2\pm 0.1)\times 10^{-2}$. The quantum numbers $J^P$ of the observed state could be found, for example, by a search for such dibaryons in the processes of charged pion photoproduction by polarized photons from the deuteron \cite{alek1}. As for the structure at 1924 MeV, the number of events in this peak exceeds the expected yields of the usual dibaryons (decaying into two nucleons) by almost two orders, even when produced with the rather big cross section of 40$\mu b/sr$. So this state, probably, could be a supernarrow dibaryon, too. It should be noted that the reaction $pd\to NX$ was investigated in other works, too (see, for example, \cite{set}). However, in contrast to the present work, the authors of these works did not study either the correlation between the parameters of the scattered nucleon and the second detected particle or the emission of the photon from the dibaryon decay. Therefore, in these works the relative contribution of the dibaryons under consideration was small, which hampered their observation. \section{Conclusions \protect\\} \label{sec:concl} The following conclusions can be made . 1) As a result of the study of the reactions $pd\to ppX_1$ and $pd\to pdX_2$ two narrow peaks at 1905 and 1924 MeV with widths less than 3 MeV have been observed in the missing mass spectra. 2) The analysis of the angular distributions of the protons and the deuterons from decay of particle $X$, produced in the reaction $pd\to pX$, showed that the peak found at 1905 MeV can be explained as a manifestation of the SND with isospin equal to 1, the decay of which into two nucleons is forbidden by the Pauli exclusion principle. 3) Probably, the observed state at 1924 MeV can be interpreted as the SND, too. \acknowledgments We thank T.E. Grigorieva and Yu.M. Burmistrov for their active participation in the creation of the setup and in the experimental runs.
\section{Introduction} High mass binary pulsars are binaries in which a radio pulsar is accompanied by an unseen companion that is approximately as massive as the pulsar. Although the companion is generally believed to be another neutron star, it appears to be likely that some of the high-mass binary pulsars are not accompanied by neutron stars but are instead accompanied by white dwarfs or low-mass main-sequence stars. PSR B1820-11 (J1823-1115), for example, is a radio pulsar in a binary with an orbital period $P\mbox{$_{\rm orb}$} \approx 357.8$~days and an eccentricity $e \approx 0.795$~(Lyne \& McKenna 1989). The derived mass function $f = 0.068\,\mbox{${\rm M}_\odot$}$ indicates that the mass of the companion exceeds $\sim 0.7\,\mbox{${\rm M}_\odot$}$. The small spin-down age of the pulsar ($\tau = 6.5$\,Myr) and its strong magnetic field ($ B \simeq 6 \times 10^{11} {\rm G}$) suggest that it is a young -- non-recycled -- pulsar (Phinney \& Verbunt 1991). Based on the mass function and the orbital eccentricity, Lyne \& McKenna (1989) suggest that the most likely companion to \mbox{PSR~B1820--11}\ is a neutron star. The companion is unlikely to be a massive hydrogen star or a black hole due to the low mass function. Lyne \& McKenna argue that the observed orbital eccentricity excludes the companion as a white dwarf or a low-mass helium star. Phinney \& Verbunt (1991) suggest that the companion of \mbox{PSR~B1820--11}\ may be a $\sim 1$\,\mbox{${\rm M}_\odot$}\ main-sequence star, which will make the system a precursor of low-mass X-ray binary. Also PSR B2303+46 may contain a non-recycled pulsar ($P = 1.06$\,s, $\tau = 30$\,Myr, $B=8 \times 10^{11}$\,G) in a binary with $P_{\rm orb} = 12.34$\,day and $e \approx 0.66$ (Stokes et al. 1985). The total mass of the binary is $2.64 \pm 0.05$\,\mbox{${\rm M}_\odot$} (Thorsett \& Chakrabarty 1999) and no optical counterpart was detected down to $R = 26^m\kern-7pt .\kern+3.5pt$ (Kulkarni 1988). Recently van Kerkwijk \& Kulkarni (1999), however, discovered an object which coincides with the timing position of \mbox{PSR~B2303+46}\ and the properties of this object are consistent with those of a massive white dwarf with a cooling age close to the age of pulsar. With population synthesis calculations we demonstrate that the birth rate of binaries in which the neutron star is born {\em after} the white dwarf is comparable to the birth rate of binaries with two neutron stars. In the former case the binary will have a high eccentricity eccentric and the radio pulsar will be accompanied by a white dwarf with a mass $\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 1.1$\,\mbox{${\rm M}_\odot$}. In this paper we argue that both \mbox{PSR~B1820--11}\ and \mbox{PSR~B2303+46}\ systems may be formed via such a scenario (see \S\,2.2). The calculations, however, do not rule out that the companion of \mbox{PSR~B1820--11}\ is either another neutron star or a main-sequence star. In the latter case, the mass of the main-sequence star would most likely be between $\sim 0.7\,\mbox{${\rm M}_\odot$}$ and $\sim 5$\mbox{${\rm M}_\odot$}. \section{The binary companions of \mbox{PSR~B1820--11}\ and PSR B2303+46} \subsection{Hiding a main-sequence star} \label{secms} At a distance of 6.3\,kpc and a height above the Galactic plane of $\sim 100$\,pc (Taylor, Manchester, \& Lyne 1993), \mbox{PSR~B1820--11}\ is heavily obscured by the interstellar medium. A star with an absolute magnitude of $M_{\rm V} \sim -4\m0$ could easily be missed in a survey with a limiting magnitude of $20\m0$ (assuming an interstellar absorption of $A_{\rm V} = 1\m6\,{\rm kpc^{-1}})$. In the USNO-A v1.0 catalog (Monet {et al.}\ \ 1997), based on the Palomar Sky Survey, the object closest to \mbox{PSR~B1820--11}\ is located at a distance of 5.9 seconds of arc and therefore cannot be associated with \mbox{PSR~B1820--11}. There is no indication for other forms of radiation (X-rays, infrared, etc.) from the direction of \mbox{PSR~B1820--11}. These suggest that the companion can not be a very massive star or have a strong stellar wind. The presence of radio emission indicates that the rate of mass loss by the companion star is modest. An upper limit to \mbox {$\dot{M}$}\ in the stellar wind of the companion can be estimated from the expression for the optical depth of the stellar wind for free-free absorption ({\it e.g.}\ , Eq.~16 in Illarionov \& Sunyaev 1975). By assuming that the wind of the accompanying star is transparent at $\lambda = 75$\,cm (and a temperature of the stellar plasma of $10^4$\,K) we obtain \begin{equation} \mbox {$\dot{M}$} \ {\raise-.5ex\hbox{$\buildrel<\over\sim$}}\ 4.8\,10^{-12} {M_{\rm tot} \over [\mbox{${\rm M}_\odot$}]} \, {P\mbox{$_{\rm orb}$} \over [{\rm day}]} \, (1-e)^{\frac{3}{2}} \,\,\; [\mbox {~${\rm M_{\odot}~yr^{-1}}$}], \label{eq:ppulsar} \end{equation} where $M_{\rm tot}$ is the mass of the binary. The estimate for \mbox {$\dot{M}$}\ depends on the eccentricity; substitution of $(1+e)$ for $(1-e)$ in Eq.~(\ref{eq:ppulsar}) gives the lower limit for the mass loss rate given that the radio pulsar is only visible at apocenter. We illustrate this limit in Fig.\, 1 for two typical combinations of $M_{\rm tot}$\ and \mbox {$\dot{M}$}. A binary with more or less similar orbital characteristics is PSR~B1259-63 ($P\mbox{$_{\rm orb}$} = 1237\pm24$\,days, $e=0.8699,\ M_{opt}= 10 \pm 3 \mbox{${\rm M}_\odot$}, \mbox {$\dot{M}$} = 5 \times 10^{-8} \mbox {~${\rm M_{\odot}~yr^{-1}}$}$; Johnston {et al.}\ 1992, 1994) which contains a radio pulsar and a Be-type star. The pulsar is visible at apastron but not at periastron. This is understood from Eq.\,(\ref{eq:ppulsar}) as an effect of shielding of the radio signal from the pulsar by the stellar wind. Another known radio pulsar in a long-period binary is PSR J0045-7319 ($P\mbox{$_{\rm orb}$} = 147.8 $\,days, $e= 0.808$) with a B1 V ($M = 8.8 \pm 1.8$\,\mbox{${\rm M}_\odot$}) companion with $\mbox {$\dot{M}$} < 3.4 \times 10^{-11} (v_{\inf}/v_{\rm esc})$\,\mbox {~${\rm M_{\odot}~yr^{-1}}$}\ (McConnell {et al.}\ 1991; Bell {et al.}\ 1995; Kaspi, Tauris \& Manchester 1996). This pulsar is visible along its entire orbit due to a more tenuous wind of the companion star, which is typical for a lower metallicity star in the Large Magellanic Cloud. The absence of eclipses and of X-rays from accretion or shocks generated by the interaction between the pulsar wind with wind of the Be star (as are observed in PSR B1259-63, Grove {et al.}\ \ 1995) also indicates that the companion of \mbox{PSR~B1820--11}\ is not likely to be a massive star. An upper limit to the mass-loss rate of the companion star in \mbox{PSR~B1820--11}\ may be set by measuring variations in the dispersion measure of the radio pulsar when it is close to periastron (see Melatos, Johnston, \& Melrose 1995). \subsection{Can the companion to a radio pulsar be a massive white dwarf?} \label{secwd} The eccentricity of a binary orbit can either be primordial or induced by the sudden mass loss by one of the stars (Blaauw 1961). The latter is expected to occur in a supernova in which the exploding star loses a considerable fraction of its mass leaving behind a neutron star. The orbital periods of \mbox{PSR~B1820--11}\ and \mbox{PSR~B2303+46}\ prohibit the observed eccentricity from being primordial; these binaries have experienced a phase of mass transfer which circularized the orbit. The orbital eccentricities of these binaries are therefore induced by a supernova which may have produced the currently observed pulsar. If the companions to \mbox{PSR~B1820--11}\ or PSR B2303+46 are white dwarfs they must have been formed {\em before} the neutron stars. This possibility was never studied in detail since the progenitor of a white dwarf is expected to live longer than the progenitor of a neutron star. However, it was noticed by Tutukov \& Yungelson (1993) and by Portegies Zwart \& Verbunt (1996) that a reversal of evolution can be accomplished if both stars are similar in mass and if the masses of both stars are only slightly smaller than the mass limit for forming a neutron star. If this is the case, the secondary star may still gain enough mass in a phase of mass transfer to pass the limit for forming a neutron star. For this to happen mass transfer should proceed rather conservatively, which will be the case since the initial mass ratio is close to unity. In the second, unstable, phase of mass transfer the white dwarf (original primary star) will spiral-in into the envelope of the secondary as this star ascends the giant branch. This phase causes the orbital period to decrease dramatically by using orbital energy to carry the common envelope to infinity. If the white dwarf and the core of the giant stay detached after the common envelope is ejected, a close binary consisting of a white dwarf and a helium star remains. If the helium star is massive enough, the binary experiences a supernova and a young radio pulsar is formed. If the system survives the supernova the companion of the radio pulsar will be a white dwarf in an eccentric orbit. The mass of the white dwarf will be $\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 1.1$\mbox{${\rm M}_\odot$} in such a case, because it originated from a star that was rather massive. \subsection{Binary population synthesis} We use the population synthesis program for binary stars {\sf SeBa}\footnote{The name {\sf SeBa} is adopted from the Egyptian word for `to teach', `the door to knowledge' or `(multiple) star'. The exact meaning depends on the hieroglyphic spelling.} (Portegies Zwart \& Verbunt 1996) for evolving a million binaries with a primary mass between 8\,\mbox{${\rm M}_\odot$}\ and 100\,\mbox{${\rm M}_\odot$}. We used model B from Portegies Zwart \& Yungelson (1998; henceforth PZY98), which satisfactorily reproduces the properties of observed high-mass binary pulsars (with neutron star companions). For a single star we adopt a minimum zero-age mass of 8\,\mbox{${\rm M}_\odot$}\ for forming a neutron stars, which coincides with estimates based on observations (Koester \& Reimers 1996, see however Ritossa, Garcia-Berro, \& Iben (1996) who show that a single star with a mass of 10\,\mbox{${\rm M}_\odot$}\ may still evolve into a oxygen-neon-magnesium white dwarf). If a star is stripped of its envelope in an early phase this lower limit may become as large as 12\,\mbox{${\rm M}_\odot$}\footnote{In {\sf SeBa} the lower limit on the progenitor mass for forming a neutron star in a binary depends on the evolutionary stage at which the star loses its hydrogen envelope due to the interaction with its companion. This depends on the orbital separation and the initial mass ratio. The lower mass limit for the formation of a neutron star in a binary may thus be higher than for single stars.}. Coincidence of the lower limits of the initial primary mass for our simulation with the minimum mass for the progenitors of neutron stars results in an underestimate in the formation of binaries in which the secondary star experiences a supernova. By comparing our results with those of Portegies Zwart \& Verbunt (1996, see their Tab.~4), who take lower mass primaries into account, we evaluate that this affects our results with $\ {\raise-.5ex\hbox{$\buildrel<\over\sim$}}\ 20$\%. While referring the reader to PZY98 for details, we list here the most crucial assumptions. The masses of the primaries obey a power-law with exponent 2.5 (Salpeter $\equiv 2.35$). The initial mass function is normalized to the current Galactic star-formation rate of 4\,\mbox{${\rm M}_\odot$}\,\mbox {{\rm yr$^{-1}$}}\ (van den Hoek \& de Jong 1997). We assume that all stars are born in binaries with a semi-major axis up to $a=10^6$~\mbox{${\rm R}_\odot$}. The distribution was chosen to be flat in $\log a$. Roche-lobe contact at zero age sets the minimum to the distribution of initial separations. The mass of the secondary is selected to be between 0.1~\mbox{${\rm M}_\odot$}\ and the mass of the primary from a flat mass ratio distribution. The eccentricity for each binary is taken from the thermal distribution between zero and unity. Neutron stars receive a kick upon birth. The velocity of the kick is taken randomly from the distribution function proposed by Hartman (1997) and in a random direction. Black holes (from stars initially more massive than 40 \mbox{${\rm M}_\odot$}) receive a kick velocity which is scaled by $M_{ns}/M_{bh}$. From computations presented in PZY98 one may infer that our conclusions concerning birth rates and \mbox{$P_{\rm orb} - e$}~\ distributions for various combinations of neutron star and companion are robust with respect to reasonable variations in the most crucial model parameters and initial conditions, such as the kick velocity distribution and common envelope parameter, initial mass function and initial distribution in orbital periods. The criteria for the stability of mass transfer and the adopted rate of mass lost via a stellar wind and the amount of momentum lost per unit mass, however, may affect the details of our calculations rather significantly. From the output of the computer simulations we select binaries with at least one neutron star. Within these constraints three orbital-period ranges are considered: $P_{\rm orb} \geq 10, 100, {\rm and}\ 1000$\ days. Table~\ref{tab:table} gives the model birth rates for the binaries in various groups depending on the nature and the mass of the companion of the neutron star. Figures 1 and 2 give the probability distributions for these binaries in the orbital period-eccentricity plane. The birth rates in Table~\ref{tab:table} reflect the initial distributions of the binary parameters and the complicated evolutionary history via several phases of mass transfer and at least one supernova. It is therefore not surprising that the most common companion to a pulsar is a rather massive ($\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 5\mbox{${\rm M}_\odot$}$)\ main-sequence star in a relatively wide binary ($ 10 \ {\raise-.5ex\hbox{$\buildrel<\over\sim$}}\ P_{\rm orb} \ {\raise-.5ex\hbox{$\buildrel<\over\sim$}}\ 1000$\,days, Fig. 1). As argued above, the presence of such a massive companion in \mbox{PSR~B1820--11}\ may be excluded by the absence of an optical counterpart and the unlikely inclination of the orbit; hiding a companion of $M\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 3\mbox{${\rm M}_\odot$}$\ demands $\cos i \ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 0.95$\ (Lyne \& McKenna 1989). The observed radio emission also makes it unlikely that the mass of the companion to \mbox{PSR~B1820--11}\ is $\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 5\mbox{${\rm M}_\odot$}$. Such a companion star with a mass loss in the stellar wind of $\ {\raise-.5ex\hbox{$\buildrel<\over\sim$}}\ 10^{-9} \mbox {~${\rm M_{\odot}~yr^{-1}}$}$ (which is chosen on the high side to account for the enhanced wind mass loss for Be stars) would easily shield the radiation of the radio pulsar along its entire orbit (see Fig.\,\ref{fig:pe1820}). In the range of orbital periods $P_{\rm orb} \geq 100$, days the birth rates of neutron stars accompanied either by another neutron star (\mbox{${\it ns}$}), a white dwarf (\mbox{${\it wd}$} ), a low mass $(M \la 1.4\,\mbox{${\rm M}_\odot$} )$~ main-sequence star ({\em ms}) or a black hole (\mbox{${\it bh}$} ) are comparable (Table \ref{tab:table}). The relatively large population of (\mbox{${\it bh}$}, \mbox{${\it ns}$}) systems and their distribution in orbital periods is a result of the evolutionary history of massive stars which is dominated by stellar wind mass loss rather than Roche lobe overflow and by the lower kick velocities imparted to black holes. However, the mass function of \mbox{PSR~B1820--11}\ makes a black hole with mass exceeding several \mbox{${\rm M}_\odot$}\ as a companion unlikely, and for \mbox{PSR~B2303+46}\ such a massive companion is excluded by the measured total mass. \begin{table}\centering \caption[]{ Results of the model computations for binaries with at least one radio pulsar. The first column identifies the various binary components. Cols.\ (2) and (3) give the mass limits for the companion of the pulsar (black holes have a mass $> 2\,\mbox{${\rm M}_\odot$}$). Col.\ (4) gives the total Galactic birth rate (BR) of such binaries. Cols.\ (5), (6) and (7) give the birth rate of binaries with an orbital period larger than 10, 100, and 1000 days, respectively, in the units of $10^{-5}$\,\mbox {{\rm yr$^{-1}$}} }. \begin{flushleft} \begin{tabular}{lrlcrrr} & $m_{\rm min}$&$m_{\rm max}$ & BR & $P_{\geq10}$ & $P_{\geq100}$ & $P_{\geq1000}$\\[5pt] & \multicolumn{2}{c}{\unit{\mbox{${\rm M}_\odot$}}} & \multicolumn{4}{c}{\unit{$10^{-5}$\mbox {{\rm yr$^{-1}$}}}} \\ [5pt] (1) & (2) & (3) & (4) & (5) & (6) & (7) \\ [5pt] (\mbox{${\it ns}$}, \mbox{${\it ms}$}) &0.1 & 0.7& 0.5 & 0.1 & 0.0 & 0.0 \\ (\mbox{${\it ns}$}, \mbox{${\it ms}$}) &0.7 & 1.4& 4.4 & 1.2 & 0.3 & 0.0 \\ (\mbox{${\it ns}$}, \mbox{${\it ms}$}) &1.4 & 5.0&13.9 & 5.0 & 2.3 & 0.3 \\ (\mbox{${\it ns}$}, \mbox{${\it ms}$}) &5.0 &10.0& 9.5 & 3.6 & 1.1 & 0.5 \\ (\mbox{${\it ns}$}, \mbox{${\it ms}$}) &10.0& up &19.7 &17.4 &10.8 & 3.3 \\ [5pt] (\mbox{${\it wd}$}, \mbox{${\it ns}$}) &1.1 & 1.4& 4.4 & 1.3 & 0.3 & 0.0 \\ (\mbox{${\it ns}$}, \mbox{${\it ns}$}) &1.4 & 1.4& 3.4 & 0.6 & 0.2 & 0.1 \\ (\mbox{${\it bh}$}, \mbox{${\it ns}$}) &2.0 & var.& 0.6 & 0.3 & 0.2 & 0.1 \\ \end{tabular} \end{flushleft} \label{tab:table} \end{table} \begin{figure} \centerline{ \hspace{-0.98cm} \psfig{file=fig1a.ps,height=5cm,angle=-90} } \vspace{-0.61cm} \centerline{\hspace{-0.8cm} \psfig{file=fig1b.ps,height=5cm,angle=-90} } \vspace{-0.61cm} \centerline{\hspace{-0.8cm} \psfig{file=fig1c.ps,height=5cm,angle=-90} } \caption[]{ Probability distribution in the orbital period -- eccentricity plane for (\mbox{${\it ns}$}, {\it ms}) binaries with a primary mass between 0.7\,\mbox{${\rm M}_\odot$}\ and 1.4\,\mbox{${\rm M}_\odot$} (upper panel), 1.4\,\mbox{${\rm M}_\odot$}\ and 5\,\mbox{${\rm M}_\odot$}\ (middle) and between 5\,\mbox{${\rm M}_\odot$}\ and 10\,\mbox{${\rm M}_\odot$}\ (lower panel, see Tab.\,\ref{tab:table} for the birth rates to which the panels are normalized). The solid line in the lower panel is plotted using Eq.\,(\ref{eq:ppulsar}) for a 5\,\mbox{${\rm M}_\odot$}\ star with a wind mass loss rate of $10^{-9}$\mbox{${\rm M}_\odot$}\,yr$^{-1}$. A binary is visible over its entire orbit to the right of both lines. To the left of both lines, the binary stays invisible throughout its entire orbit. In between the left and the right part of the solid lines the binary is visible for part of its orbit (near apocenter). The dashed line is plotted using a 10\,\mbox{${\rm M}_\odot$}\ primary with a mass loss rate of $10^{-7}$\mbox{${\rm M}_\odot$}\,yr$^{-1}$. The $\star$ symbol indicates the position of PSR~B1820-11, and two bullets (lower panel) indicate the positions of PSR~B1259-63 (right) and J0045-7319 (left).} \label{fig:pe1820} \end{figure} \begin{figure} \centerline{\hspace{-0.8cm} \psfig{file=fig2a.eps,height=5cm,angle=-90} } \vspace{-0.61cm} \centerline{\hspace{-0.8cm} \psfig{file=fig2b.eps,height=5cm,angle=-90} } \caption{Probability distributions for binaries which contain a young neutron star and a white dwarf (top panel) or a neutron star (lower panel) as a companion star. A figure for (\mbox{${\it bh}$}, \mbox{${\it ns}$}) is not provided because of the small number of data points. The $\star$\ symbols identify the positions of \mbox{PSR~B1820--11}\ (right) and \mbox{PSR~B2303+46}\ (left). Dots in the lower panel give the positions of the known recycled pulsars with a neutron star companion. } \label{figallpe} \end{figure} Figure~\ref{figallpe} compares the probability distributions for binaries with a radio pulsar which is accompanied by a white dwarf (upper panel) and an older neutron star (lower panel) in the orbital period - eccentricity diagram. Our model calculations reveal that the binaries in which a young pulsar is accompanied by a massive white dwarf are generally rather eccentric ($e \sim 0.6$) and that their orbital periods are distributed with more or less equal probability in $\log P_{\rm orb}$ between a few hours and about one hundred days. In Fig.\, \ref{fig:pe1820} \mbox{PSR~B1820--11}\ appears at the edge of the area with highest probability, while the orbital parameters of \mbox{PSR~B2303+46}\ fall nicely in the densely populated area of the \mbox{$P_{\rm orb} - e$}~\ diagram. This suggests, together with the rather high birth rate of such binaries, that \mbox{PSR~B2303+46}\ may well contain a white dwarf as a companion star. Recently van Kerkwijk \& Kulkarni (1999) found evidence that \mbox{PSR~B2303+46}\ is accompanied by a white dwarf with a mass of about 1.2\,\mbox{${\rm M}_\odot$}. \mbox{PSR~B1820--11}\, fits rather ill in either of the \mbox{$P_{\rm orb} - e$}~\ diagrams for (\mbox{${\it ns}$}, \mbox{${\it ns}$}), (\mbox{${\it wd}$}, \mbox{${\it ns}$}), and low mass or moderate mass (\mbox{${\it ns}$}, \mbox{${\it ms}$}) binaries. The relative birthrates for a population of binaries with long orbital periods increases by increasing the efficiency of the deposition of orbital energy into common envelopes during the spiral-in phase (see PZY98). However, this would require us to relax the decisive criterion why we chose this specific set of model parameters; the coincidence of having the other high mass binary pulsars (PSR J1518+49, B1534+12, and B1913+16) in the most densely populated areas in the \mbox{$P_{\rm orb} - e$}~\ diagram for (\mbox{${\it ns}$}, \mbox{${\it ns}$}) binaries. There is no evidence for a strong selection against the discovery of a young pulsar which is accompanied by a dead pulsar or by a massive white dwarf with an orbital period $\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 10$\,days (Johnston \& Kulkarni 1991). The higher birth rate of binaries in which the radio pulsar is accompanied by a low mass main-sequence star and the better coincidence in the \mbox{$P_{\rm orb} - e$}~\ diagram makes this possibility very attractive. If the companion of \mbox{PSR~B1820--11}\ is indeed a low-- or moderate mass main-sequence star it may be a precursor of a system similar to GX~1+4 = V2116 Oph, which shows the X-ray features of an accreting neutron star and which is identified with a symbiotic star (Davidsen, Malina, \& Bowyer 1977). The orbital period of GX~1+4 is $\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 100$\, days (Chakrabarty \& Roche 1997). The low birth rate of such binaries (Table\, \ref{tab:table}) together with the short life time of a red giant explains the paucity of such X-ray binaries. \section{Conclusion} We have studied the nature of the companions of young radio pulsars in eccentric orbits. Our calculations reveal that the possibility that a young pulsar is accompanied by a white dwarf of mass $\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 1.1\mbox{${\rm M}_\odot$}$ cannot be neglected. The birth rate of such binaries is higher than the birth rate of of binaries in which a young pulsar is accompanied by a dead pulsar (i.e.: an old neutron star). We argue therefore that \mbox{PSR~B1820--11}\ may be accompanied by a massive white dwarf, however the probability that its companion is a main-sequence star with a mass $\ {\raise-.5ex\hbox{$\buildrel<\over\sim$}}\ 5\mbox{${\rm M}_\odot$}$ (or even another neutron star) is similar to the probability of being a white dwarf. Our calculations confirm the observation of van Kerkwijk \& Kulkarni (1999) in the sense that it is likely that \mbox{PSR~B2303+46}\ is accompanied by a white dwarf with a mass $\ {\raise-.5ex\hbox{$\buildrel>\over\sim$}}\ 1.2\,\mbox{${\rm M}_\odot$}$. \section*{Acknowledgments} We thank Thereasa Brainerd, Ramach Ramachandran and the anonymous referee for reading the manuscript. We acknowledge the kind hospitality of the Astronomical Institute 'Anton Pannekoek', Meudon Observatory and the Univeristy of Tokyo where a part of this work was done. This work is supported by NWO Spinoza grant 08-0 to E.~P.~J.~van den Heuvel, RFBR Grant 96-02-16351, the JSPS fellowship to SPZ and by NASA through grnat number HF-01112.01-98A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS\, 5-26555.
\section{Introduction} Tunneling is one of the most striking predictions of quantum mechanics, and continues to provoke some of its most heated controversies. Despite the appearance of tunneling phenomena in numerous physical and technological areas (and in first-year physics courses), certain aspects of the effect remain poorly understood. This is seen most clearly in the debate over how long a particle takes to traverse a tunnel barrier, and in particular, whether or not it can do so faster than light \cite{Chiao=1997ProgOpt,Steinberg=1998AnnPhys}. The confusion over these issues can be traced to certain common elements of quantum ``paradoxes.'' For one, definite trajectories cannot be assigned to particles in general, and in this sense it is not even clear how to rigorously phrase a question about how much time a transmitted particle spent in a forbidden region-- in fact, it may not even be necessary that a particle ``traverse'' a region in order to be found on the far side. Of course, at one level quantum mechanics is merely a wave theory, and quite thoroughly understood. In many physical situations, more controversial, interpretational, issues (related to ``collapse'' or other alternatives) may easily be skirted without loss of predictive power. In tunneling, however, it is quite natural to look for a description of transmitted particles, as distinct from reflected ones (or from the ensemble as a whole) \cite{Buttiker=1982}. But such a description is impossible without an attempt to model the detection itself, because without the detection event, transmitted and reflected packets necessarily coexist. Detection naturally raises other interesting questions. What is the nature of a detection process which occurs {\it inside} a forbidden region (cf. \cite{Carniglia=1972JOSA,Carniglia=1971PRD})? According to the collapse postulate, if a particle is found to be in the barrier region, it is subsequently described by a new wave function, confined to that region. Any such wave function has $E>V_0$, and suddenly, the problem is no longer one of tunneling. Our plans to observe tunneling of laser-cooled Rubidium atoms, and to perform ``weak measurements''\cite{Aharonov=1990,Aharonov=1988} in order to study the behaviour of a transmitted subensemble, have been presented at length elsewhere\cite{Steinberg=1998Superlatt,Steinberg=1998AnnPhys,Myrskog=1999}. Here we repeat only the essential elements, to provide context for the present discussion. \section{Tunneling in atom optics} Laser-cooled atoms offer a unique tool for studying quantum phenomena such as tunneling through spatial barriers. They can routinely be cooled into the quantum regime, where their de Broglie wavelengths are on the order of microns, and their time evolution takes place in the millisecond regime. They can be directly imaged, and if they are made to impinge on a laser-induced tunnel barrier, transmitted and reflected clouds should be spatially resolvable. With various internal degrees of freedom (hyperfine structure as well as Zeeman sublevels), they offer a great deal of flexibility for studying the various interaction times and nonlocality-related issues. In addition, extensions to dissipative interactions and questions related to irreversible measurements and the quantum-classical boundary are easy to envision.\cite{Steinberg=1998Superlatt} In our work, we prepare a sample of laser-cooled Rubidium atoms in a MOT, and cool them in optical molasses to approximately 6 $\mu K$. As explained below, further cooling techniques are under investigation for achieving yet lower temperatures \cite{Myrskog=1999}. We plan to use a tightly focussed beam of intense light detuned far to the blue of the D2 line to create a dipole-force potential for the atoms\cite{Rolston=1992,Miller=1993FORT,Davidson=1995}. In this intense beam, the atom becomes polarised, and the polarisation lags the field by $90^{\circ}$ when the light frequency exceeds that of the atomic resonance. This polarisation out of phase with the local electric field constitutes an effective repulsive potential, proportional to the intensity of the perturbing light beam. It can also be thought of in terms of the new (position-dependent) energy levels of the atom {\it dressed} by the intense laser field. Using a 500 mW laser at 770 nm, we will be able to make repulsive potentials with maxima on the order of the Doppler temperature of the Rubidium vapour. Acousto-optical modulation of the beam will let us shape these potentials with nearly total freedom, such that we can have the atoms impinge on a thin plane of repulsive light, whose width would be on the order of the cold atoms' de Broglie wavelength. This is because the beam may be focussed down to a spot several microns across (somewhat larger than the wavelength of atoms in a MOT, but of the order of that of atoms just below the recoil temperature, and hence accessible by a combination of cooling and selection techniques). This focus may be rapidly displaced by using acousto-optic modulators and motorized mirrors. As the atomic motion is in the mm/sec range, the atoms respond only to the time-averaged intensity, which can be arranged to have a nearly arbitrary profile. As a second stage of cooling, we follow the MOT and optical molasses with an improved variant of a proposal termed ``delta-kick cooling''\cite{Ammann=1997}. In our version, the millimetre-sized cloud is allowed to expand for about ten milliseconds, to several times its initial size. This allows individual atoms' positions $x_{i}$ to become strongly correlated with atomic velocity, $x_{i} \approx v_{i}t_{\rm free}$. Magnetic field coils are then used in either a quadrupole or a harmonic configuration to provide a position-dependent restoring force for a short period of time. By proper choice of this impulse, one can greatly reduce the rms velocity of the atoms. So far, we have achieved a one-dimensional temperature of about 700 nK, corresponding to a de Broglie wavelength of about half a micron. We are currently working on improving this temperature by producing stronger, more harmonic potentials, and simultaneously providing an antigravity potential in order to increase the interaction time. However, the tunneling probability through a 5-micron focus will still be negligible at these temperatures. Furthermore, the exponential dependence of the tunneling rate on barrier height will be difficult to distinguish from the exponential tail of a thermal distribution at high energies. We will therefore follow the delta-kick with a velocity-selection phase\cite{Myrskog=1999}. By using the same beam which is to form a tunnel barrier, but increasing the width to many microns, we will be able to ``sweep'' the lowest-energy atoms from the center of the magnetic trap off to the side, leaving the hotter atoms behind. Our simulations suggest that we will be able to to transfer about 7\% of the atoms into the one-dimensional ground state of this auxiliary trap. This new, smaller sample will have a thermal de Broglie wavelength of approximately $3.5 \mu$m, leading to a significant tunneling probability through a 10-micron barrier. We expect rates on the order of 1\% per secular period, causing the auxiliary trap to decay via tunneling on a timescale of the order of 100 ms. \section{Measuring tunneling atoms} A weak measurement is one which does not significantly disturb the particle being studied (nor, consequently, does it provide much information on any single occasion). Why not perform a strong measurement? Simply because if one can tell with certainty that a particle is in a given region, one has also determined that the particle has enough energy to be in that region; one is no longer studying tunneling. The measurement has too strongly disturbed the unitary evolution of the wave function. At the 6th Symposium on Laser Spectroscopy in Taejon, I made the above glib assertion as I had frequently done in the past, and went on to discuss weak measurements. Afterwards, however, Bill Phillips raised the question of {\it where} exactly the energy comes from to turn a forbidden region into an allowed one. The imaging of an atom involves a small transfer of momentum, and typically the only energy exchange is the more-or-less negligible recoil shift. But in this scenario, an atom observed under an arbitrarily high tunnel barrier must-- merely by being observed-- acquire enough energy to ride on top of the barrier. Why should the effect of a weak probe beam (in fact, the interaction with a single resonant photon) scale with the completely unrelated height of a potential barrier? The situtation envisioned is shown in schematic form in Fig. 1. The wave function of the atom decays exponentially into the barrier region over a characteristic length $1/\kappa$. If this length is greater than the resolution of the imaging lens, then it is possible for the appearance of a spot of focussed fluorescence on an appropriate point on the screen to indicate that an atom is in the barrier region. This atom, having scattered perhaps only a single photon, must according to quantum mechanics have acquired an energy of at least $V_0$ to exist confined to the barrier region. This energy depends not on the wavelength or intensity of the imaging light, but only on the height of the barrier created by the dipole-force beam, which may greatly exceed the recoil energy associated with the momentum transfer involved in elastic scattering of a photon. An interesting point is that it is unnecessary to actually focus and detect the photon in question. The very possibility that some future observer {\it could} use the scattered light to determine that an atom was in the forbidden region is sufficient to decohere spatially separated portions of the atomic wavefunction, causing some fraction of the atoms to ``collapse'' (if you will) into the barrier region. At first, one might think that the energy comes from the interaction between the atom and the dipole-force beam. A little thought suffices, however, to demonstrate that this cannot be the solution. Even in the absence of a potential, an imaging beam may localize a previously unlocalized particle, increasing its momentum uncertainty and hence its energy. The energy must come from the imaging beam. Why, then, does the quantity of energy transferred depend on the barrier height? A partial answer comes from carefully considering the energy levels of the atom. Inside the barrier region, the presence of the dipole beam couples the atomic eigenstates, creating an AC Stark shift (which is the effective repulsive potential). An atom which makes a transition between a state primarily outside the barrier (of energy $E_g + P^2/2m$) to a state localized in the barrier is simultaneously making a transition to a new, higher-energy electronic state ($E_{g}+V_{0}+P^{\prime\, 2}/2m$). Energy-conservation will be enforced by the time-integral in perturbation theory, causing the amplitudes for this process to interfere destructively unless the scattered photon energy plus the final energy of the atom equals the initial photon energy plus the initial energy of the atom. In other words, the presence of the dipole beam makes possible inelastic (Raman) transitions between different atomic states. When an elastic scattering event occurs, the atom is left in the original state, and cannot be localized to the barrier. Only when an inelastic collision occurs can the atom be transferred to the state dressed by the dipole field, and localized in the formerly forbidden region. Can one then determine that an atom is in the barrier {\it without} imaging, by merely measuring the energy of the scattered photon? Unfortunately, no. Recall that this argument hinges on an imaging resolution \begin{equation} \delta l < 1/\kappa \;, \end{equation} where \begin{equation} \hbar^2\kappa^2 = 2m(V_0-E) \; . \end{equation} A particle localized to within $\delta l$ has a momentum uncertainty \begin{equation} \Delta P \geq \hbar/2\delta l\; . \end{equation} This means that it will only remain within the resolution volume for a time \begin{equation} t \leq \frac{\delta l}{\Delta P/m} = 2m\delta l^2/\hbar\; . \end{equation} This in turn implies \begin{equation} t < 2m/\hbar\kappa^2\; . \end{equation} Unless the imaging light is time-resolved to better than this limit, it is impossible to maintain the spatial resolution necessary to conclude with certainty that the particle is in the barrier. (Strictly speaking, it would suffice to image to better than the barrier width. However, a particle in the barrier is most likely to be within the first exponential decay length $1/\kappa$. While with lower resolution, one might still conclude that the particle was deeper in the barrier, the likelihood will be exponentially suppressed. Thus on some occasions, the photon energy will be shifted by an amount greater than its rms spectrum, but the low amplitude of this frequency component will be matched by the low probability of finding the atom so deep in the barrier, and the present arguments may easily be generalized.) This implies that the energy uncertainty of the scattered photon must be \begin{equation} \Delta E \stackrel{>}{_\sim} \hbar/t > \hbar^2\kappa^2/2m \; . \end{equation} But this is $V_0-E$, just the energy required to excite the tunneling atom above the barrier. So the only way to image an atom in the forbidden region is to use light with sufficient energy uncertainty that it can boost the atom above the barrier {\it without} a significant change in its own spectrum. \section{Detection without observation, or observation without detection?} Just as these issues were beginning to make themselves clear to us, Terry Rudolph suggested an even more confounding extension. His idea is outlined in Figure 2. Suppose one decides to determine the location of the tunneling atom in a more indirect manner. Specifically, suppose a nearly ideal imaging system is devised (relying, for example, on $\pi$-pulses of probe light), but that a beam stop is imaged onto the barrier region. In this way, any atom in the classically allowed region will be imaged, but an atom which finds itself in the forbidden region will be out of the reach of probe light, and no photon will be scattered. When no scattered photon is observed, we can conclude with near certainty that the atom is in the forbidden region, and has therefore made a transition to a higher-energy dressed state. But now where did the energy come from? After all, it appears that the ``detected'' atom became localized without ever undergoing an interaction. This picture is ill-founded, however. The atom cannot be considered in isolation; it is in fact the entire system, composed of atom, dipole-force beams, and imaging photons, which undergoes a transition and must conserve energy. Under the influence of a probe pulse, the atom's quantum state becomes entangled with the state of the imaging light. There is some amplitude for a photon to travel along its original path, unscattered, but this amplitude is correlated with an atomic state localized to the barrier region. For the time-integral of this amplitude to lead to a real probability for detecting an unscattered photon, the detected photon will necessarily lose enough energy to boost the atom above the barrier, just as in the case previously discussed. Once more, the situation becomes less startling when we observe that (1) there is indeed a mechanism for energy-exchange between the ({\it unscattered}) imaging beam and the atom; and (2) this energy exchange never exceeds the intrinsic uncertainty in the initial photon energy. The interaction between imaging light and an atom comprises not only the possibility of scattering, but also the real part of the atomic polarizability, which is to say the index of refraction experienced by the light due to the presence of the atom. For a near-resonant photon with a probability $\eta$ of being scattered by an atom, the extra optical path introduced by the presence of the atom, $\int [n(z)-1]dz$, is of the order of an optical wavelength times $\eta$, corresponding to an optical phase shift approximately equal to $\eta$. If an atom is found to have appeared in the dark region enclosing the barrier, this implies that it left the region of interaction with the probe light, causing the light to experience a time-varying index of refraction. If the atom's departure from the illuminated region is known to have occurred within a time $t$, then the phase of the light was modulated by an amount $\eta$ in a time smaller than $t$, producing a frequency shift of the order $\eta/t$. Each photon's energy can in this way be altered by the ``disappearance'' of the atom, by the quantity $\hbar\eta/t$. Since on the order of $1/\eta$ photons are necessary to detect atoms with near-unit probability in such a scenario, this phase-modulation effect is automatically sufficient to produce an energy exchange of up to $\hbar/t$ between the moving atom and the probe beam, {\it even when no photons are scattered}. As in the original discussion of {\it bright} imaging of the barrier region, we know that $\hbar/t \stackrel{>}{_\sim} \hbar^{2}\kappa^{2}/2m$, and this energy exchange is enough to propel the particle above the barrier. Furthermore, the same argument concerning the duration of a probe pulse remains intact. If the pulse lasts long enough that even a particle localized to the barrier would have time to escape while the light was on, then one will never completely avoid fluorescence, and thus never be able to conclude with certainty that the particle is in the barrier region. One might instead envision a CW probe but time-gated photodetection; in this case, the argument is similar, but it is the {\it detected} photon whose energy can no longer be determined precisely enough to be certain that energy exchange has taken place on any individual occasion. Nevertheless, by studying an ensemble of particles, one should be able to build up enough statistics to confirm the shift in the mean photon frequency. \section{Conclusion} We see that tunneling is just one more prototypical example of the way in which observation may disturb a quantum system. It is instructive to consider the mechanisms which allow the necessary energy transfer to take place, along with the requisite uncertainties behind which this transfer hides. Ultracold atoms in Bose condensates, and at temperatures achievable through related laser-cooling techniques as well, have long enough de Broglie wavelengths that tunneling effects should soon be observed in a regime where these questions become more than purely academic. Particularly intriguing is the possibility of {\it modifying} the barrier-traversal rate by the application of a probe beam which {\it could} in principle be used to image an atom in the forbidden region. Even if no attempt is made to actually perform the imaging, the simple possibility that one could do so should be enough to turn the quantum amplitude for an atom to be within about $1/\kappa$ of the edge of the barrier into an actual probability, in the sense of a real fraction of atoms localized into that region of the barrier. These atoms have enough energy to traverse the barrier classically in either direction, and may therefore be observed on the far side. \section{Acknowledgments} This discussion would remain purely academic if not for the hard work of Stefan Myrskog, Jalani Fox, Phillip Hadley, and Ana Jofre on our laser-cooling experiment. I would also like to acknowledge Jung Bog Kim and his students Han Seb Moon and Hyun Ah Kim for their collaboration on this project. I want to thank Jung Bog Kim and the organizers of the Symposium on Laser Spectroscopy for their invitation and for their hospitality during the meeting, which proved quite stimulating. Finally, I would like to thank Bill Phillips for the question which prompted this short paper, and for fascinating discussions concerning it; and Terry Rudolph for following it up with an even harder question just as I thought I was beginning to understand something.
\section{Abstract} A sample of 323 Ultraluminous IRAS galaxies (ULIRGs) has been correlated with the ROSAT All-Sky Survey and ROSAT public pointed observations. 22 objects are detected in ROSAT survey observations, and 6 ULIRGs are detected in addition in ROSAT public pointed observations. The detection is based on a visual inspection of the X-ray contour maps overlaid on optical images of ULIRGs taken from the Digitized Sky Survey. Simple power law fits were used to compute the absorption-corrected fluxes of the ROSAT detected ULIRGs. The ratio of the soft X-ray flux to the far-infrared luminosity is used to estimate the contribution from starburst and AGN emitting processes. These results are compared with the ISO SWS ULIRG diagnostic diagram. \section{Class properties of ULIRGs} \subsection{Specific observational results} \subsubsection{IRAS 10026+4347} The ULIRG IRAS 10026+4347 can also be classified as a narrow-line quasar. The FWHM of the $\rm H\beta$ line is about 2500 $\rm km\ s^{-1}$, and strong optical Fe II multiplet emission is a prominent feature of its optical spectrum. The X-ray spectrum exhibits a steep X-ray continuum slope, with a photon index for a simple power law fit of $\rm \Gamma = 3.2 \pm 0.5$, typical of narrow-line Seyfert 1 galaxies and narrow-line quasars. The (0.1$-$2.4 keV) luminosity of IRAS 10026+4347, obtained via a simple power law fit to the data, and corrected for absorption by neutral hydrogen along the line of sight, is $\rm 1.12 \cdot 10^{45}\ erg \ s^{-1}$. The ratio of the soft X-ray (0.1$-$2.4 keV) to far-infrared (40$-$120 $\rm \mu$m) flux is 0.25. In Section 2.3 we argue that values of this ratio above about 0.003 require a contribution of an AGN component, in addition to starburst processes. \subsubsection{Mrk 231} Mrk 231 (IRAS 12540+5708) is detected both in the ROSAT All-Sky Survey and in ROSAT public pointed observations. The X-ray light curve (cf. Fig. 1), obtained from a ROSAT pointed observation, suggests some indication of variability with a doubling time scale of about 0.4 days. \subsection{ROSAT-detected ULIRGs} 22 of the 323 ULIRGs from the IRAS 1.2 Jansky redshift catalogue (Fisher et al. 1995) are detected in the ROSAT All-Sky Survey (cf. Table 1). By inspecting the structure of the X-ray emission in overlays on optical images taken from the Digitized Sky Survey, it is strongly believed, that the objects in Table 1 are potential identifications of ULIRGs in soft X-rays. Table 2 lists the ULIRGs detected in public ROSAT pointed observations, in addition to the objects detected in the ROSAT All-Sky Survey. 6 objects are detected in pointed observations, resulting in a total number of 28 ULIRGs detected with ROSAT. Although the ROSAT energy range does not allow the probing of highly obscured regions in ULIRGs, the ROSAT All-Sky Survey allows at least a statistical approach to the class properties of ULIRGs. In Section 2.3 we discuss the ratio of the soft X-ray to far-infrared luminosity of ROSAT detected ULIRGs, which can be used to estimate the relative fraction of starburst emitting processes and emission due to accretion onto supermassive black holes. \renewcommand{\baselinestretch}{0.8} \begin{figure} \centerline{\psfig{figure=boller_fig3.ps,width=12.0cm,height=6.0cm,clip=}} \caption{ {\scriptsize ROSAT PSPC light curve of the ULIRG Mrk 231. The count rate increase suggests intrinsic variability in the ULIRG Mrk 231. The estimated upper limit for the size of the emitting region is about $\rm 10^{15}\ cm$. } } \end{figure} \renewcommand{\baselinestretch}{1.0} \renewcommand{\baselinestretch}{0.8} \begin{figure} \mbox{ \psfig{figure=plot_fxfir_new.ps,width=8.3cm,height=9.0cm,clip=} \psfig{figure=plot_fxfir_2.ps,width=8.3cm,height=9.5cm,clip=}} \caption{ {\scriptsize {\bf Left Figure:} Ratio of the 0.1$-$2.4 keV soft X-ray flux to the far-infrared (40$-$120 $\rm \mu$m) flux versus the far-infrared flux for ROSAT detected ULIRGs. The $\rm f_X/f_{IR}$ ratio covers 4 orders of magnitude. {\bf Right Figure:} Ratio of the soft X-ray luminosity to the far-infrared luminosity versus the far-infrared luminosity for ROSAT-detected ULIRGs. The dashed lines give lines of constant X-ray luminosity. The evolutionary tracks were obtained from the population syntheses of star-formation processes with star-formation rates of 50 and 500 solar masses per year. The mean value of the soft X-ray to far-infrared luminosity emitted by star-formation processes is about 0.003. Assuming a time-variable star-formation rate, with an increase of the star-formation rate by a factor of 10 over $\rm 10^8$ years, the ratio between the soft X-ray and the far-infrared luminosity varies by about a factor of 3. This suggests that objects with ratios above about 0.003 are powered by accretion processes onto a supermassive black hole, in addition to star-formation processes. } } \end{figure} \renewcommand{\baselinestretch}{1.0} \begin{table} \begin{center} \begin{tabular}{|llll|} \hline (1) & (2) & (3) & (4) \\ IRAS (NED) name & log & log & log \\ & $\rm L_{FIR}$ & $\rm L_X$ & $\rm \frac{L_X}{L_{FIR}}$\\ & [erg/s] & [erg/s] & \\ 01166$-$0844 (-) & 45.65 &42.89& -2.79\\ 01268$-$5436 (-) & 45.49 &44.53& -0.99\\ 01572+0009 (PG 0157+001) & 46.05 &44.69& -1.40\\ 02483+4302 (-) & 45.39 &43.15& -2.25\\ 04454$-$4838 (ESO203-IG1) & 45.48 &42.57& -2.92\\ 05189$-$2524 (-) & 45.66 &42.51& -3.16\\ 05494+6058 (-) & 45.43 &43.96& -1.50\\ 11257+5850 (NGC 3690) & 45.34 &41.74& -3.60\\ 11598$-$0112 (-) & 46.04 &44.42& -1.66\\ 12265+0219 (3C 273) & 46.05 &46.06& -0.04\\ 14394+5332 (-) & 45.64 &43.59& -2.08\\ 15223$-$3604 (-) & 45.46 &43.75& -1.74\\ 15462$-$0450 (-) & 45.76 &43.66& -2.13\\ 16504+0228 (NGC 6240) & 45.44 &42.84& -2.61\\ 22482$-$7027 (-) & 45.62 &41.84& -3.81\\ 22537$-$6511 (PKS 2253-65) & 45.59 &43.32& -2.30\\ 23410+0228 (-) & 45.56 &43.27& -2.31\\ 15327+2340 (Arp 220) & 45.83 &41.78& -4.04\\ 13428+5608 (MRK 273) & 45.77 &42.05& -3.73\\ 12540+5708 (Mrk 231) & 46.03 &42.19& -3.85\\ 09320+6134 (UGC 5101) & 45.62 &41.54& -4.09\\ 10026+4347 (-) & 45.59 &45.05& -0.60\\ \hline \end{tabular} \end{center} \caption{ {\scriptsize ULIRGs detected in the ROSAT All-Sky Survey. Column 1 lists the name from the IRAS Point Source Catalogue, and when appropriate, the NED name. The far-infrared luminosity and the soft X-ray luminosity are listed in columns 2 and 3, respectively. The last column gives the ratio of the soft X-ray to far-infrared luminosity. This ratio covers about 4 orders of magnitude. } } \end{table} \begin{table} \begin{center} \begin{tabular}{|llll|} \hline (1) & (2) & (3) & (4) \\ IRAS (NED) name & log & log & log \\ & $\rm L_{FIR}$ & $\rm L_X$& $\rm \frac{L_X}{L_{FIR}}$\\ & [erg/s] & [erg/s] & \\ 07598+6508 &45.86 &43.37 &-2.53\\ ([HB89]0759+651) & & & \\ 13451+1232 (-) &45.73 &42.69 &-3.08\\ 14348$-$1447 (-) &45.95 &42.56 &-3.42\\ 15033$-$4333 (-) &45.73 &42.67 &-3.09\\ 15250+3609 (-) &45.61 &42.02 &-3.60\\ 20551$-$4250 (JB40) &45.63 &42.05 &-3.59\\ \hline \end{tabular} \end{center} \caption{ {\scriptsize ULIRGs detected in public ROSAT pointed observations. Objects which are detected in the ROSAT All-Sky Survey are not included in this table. For the description of the columns see Table 1. } } \end{table} \subsection{The soft X-ray to far-infrared flux ratio} Since the total X-ray luminosity of a star-forming galaxy is proportional to its total star-formation rate, one might assume that a high X-ray luminosity might just reflect a high star-formation rate. A problem with this picture arises when one compares the soft X-ray (0.1$-$2.4 keV) flux with the far-infrared (40$-$120 $\rm \mu$m) flux. Both quantities are proportional to the star-formation rate, and an increase in the star-formation rate results in the first order in a horizontal shift of an object in Fig. 2. We (Boller \& Bertoldi 1996) found that in equilibrium, the ratio of the soft X-ray to far-infrared flux is about 0.003. Considering variable star-formation rates, the ratio between both quantities varies by about a factor of 3 (see the evolutionary tracks in the right panel of Fig. 2, where an increase of the star-formation rate by a factor of 10 for a time scale of $\rm 10^8$ years is assumed). The total far-infrared fluxes were computed from the IRAS 60 and 100 $\rm \mu$m fluxes following Helou (1985). To compute the soft X-ray fluxes from the PSPC count rate, a simple power law spectrum of the form $\rm E^{-\alpha}$ was assumed. The fluxes were converted into luminosities using eq. 7 of Schmidt \& Green (1986). A Hubble constant of $\rm H_0 = 50\ km\ s^{-1}\ Mpc^{-1}$ and a cosmological deceleration parameter of $\rm q_0 = 0.5$ were adopted. In Fig. 2 (left panel) the ratio between the soft X-ray and far-infrared flux is plotted against the far-infrared flux. The right panel gives the corresponding distribution for luminosities. The ratio between the soft X-ray and far-infrared flux ranges over 4 orders of magnitude. An additional AGN contribution is necessary to reach the X-ray to far-infrared ratio for those objects with values above 0.003. \section{Comparison with ISO SWS results} \renewcommand{\baselinestretch}{0.8} \begin{figure} \centerline{\psfig{figure=plot_pha.ps,width=11cm,height=9cm,clip=}} \caption{ {\scriptsize ROSAT-detected ULIRGs with known [O IV], Ne II and PAH feature measurements in the ISO SWS diagnostic diagram. The sizes of the circles scale with the soft X-ray to far-infrared flux ratio. All ROSAT-detected ULIRGs in the diagram are predominantly powered by star-formation processes (see Section 2.3). The ROSAT-detected ULIRGs are located in that region of the ISO SWS diagram where the AGN contribution is less than 50 per cent. } } \end{figure} \renewcommand{\baselinestretch}{1.0} In Fig. 3 the ROSAT results are compared with the diagnostic diagram obtained from ISO SWS measurements to distinguish between starburst and AGN processes in ULIRGs. The ISO SWS diagnostic diagram shows the ratio between the high- and low-excitation fine structure lines versus the strength of the PAH $\rm 7.7\ \mu$m feature (see Lutz, this proceedings). The circles indicate ROSAT-detected ULIRGs which have measured values in the ISO diagnostic diagram. The size of the circle scales with increasing soft X-ray to far-infrared flux ratio. All ROSAT detected-ULIRGs in this diagram are predominantly powered by star-formation processes, as the ratio for all objects is below the critical value of 0.003. This is in agreement with the prediction from the ISO measurements, as all objects are located in Fig. 3 in the region where the AGN contribution is less than 50 per cent. For ROSAT-detected ULIRGs with ratios of the soft X-ray to far-infrared flux above a value of 0.003, no ISO SWS [O IV], Ne II or PAH feature measurements are available. \section{Future prospects for the study of ULIRGs} We have proposed to observe the most interesting ULIRGs within the guaranteed time program of XMM. We intend to extend our studies by precisely determining the spectral and timing properties of ULIRGs, to further disentangle starburst- and AGN-emitting processes in ULIRGs.
\section{Higgs Couplings and masses: Theoretical predictions} \subsection{Predictions in the SM:}\label{twoone} In the SM the existence of Higgs boson is necessary to bring about the SSB which gives masses to the fermions and the gauge bosons, still keeping the theory renormalisable. For the SSB to happen, the $\mathrm{(mass)}^2$ term for the complex scalar doublet $\Phi$ has to be negative, i.e. the potential $V(\Phi)$ is \begin{equation} V(\Phi) = {\frac {\lambda} {4!}} (\Phi^\dag \Phi)^2 - \mu^2 \Phi^\dag \Phi \label{e1p} \end{equation} with $\mu^2$ positive. After the SSB, out of the four scalar fields which comprise $\Phi$, we are left with only the physical scalar $h$ with a mass \begin{equation} m_h^2 = \lambda \mathit{v}^2 \label{e1} \end{equation} Further, the tree level couplings of the Higgs boson $h$ to the SM fermions and the gauge bosons are uniquely determined and are proportional to their masses. The coupling of a Higgs to a pair of gluons/photons does not exist at the tree level, but is induced at one loop level by the diagrams shown in \begin{figure}[htb] \begin{center} \mbox{\epsfig{file=triangle.eps,height=30mm}} \end{center} \caption{Loop diagrams responsible for $h \rightarrow \mbox{$\gamma \gamma$} (\mbox{gg})$. \label{figure1}} \end{figure} Fig.~\ref{figure1}. As with the other couplings, this coupling is also completely calculable given the particle content of the SM, to a given order in the strong and electromagnetic coupling~\cite{spira}. The $h$gg coupling is dominated by the top quark contribution in the loop diagram whereas the $h$\mbox{$\gamma \gamma$}\ coupling receives dominant contribution from both, the top loop as well as the $W$ loop. These channels have appreciable branching ratios, albeit very small, only for $m_h \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 2 m_W$. Recall also that the precision measurements at the $Z$ indicate $\mbox{$m_h$} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 300$ GeV~\cite{LEPCH}. It should be noted that the QCD corrections for $\Gamma (h \rightarrow \mbox{gg})$ are significant $(\sim 65 \%)$~\cite{spira}. In the intermediate mass range (i.e. $m_h \leq 140$ GeV) the total width of the Higgs is $\:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 10 $ MeV, dominant decay is into a $b \bar b$ final state (e.g. for $m_h = 120 $ GeV $\Gamma (h \rightarrow b \bar b) \simeq 68 \% )$; on the other hand the branching ratio into a \mbox{$\gamma \gamma$}\ final state is about one part per mille. The total width is $\sim 1 $ GeV around $m_h \sim 300 $ GeV and rises very fast after that, reaching $\Gamma_h \sim m_h$ around $m_h \:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: $ 500 GeV. Calculations of various branching ratios, including higher order QCD effects are available~\cite{spira}. While the various couplings and hence the branching fractions of the Higgs are well determined once \mbox{$m_h$}\ and various other parameters in the SM such as $m_t, \alpha_s$ etc. are specified, \mbox{$m_h$}\ itself is completely undetermined in the SM. However, as seen from Eq.~\ref{e1}, it is linearly related to the self coupling of the scalar field. Eventhough, the theory has nothing to say about the Higgs mass {\it per se} the behaviour of the self coupling $\lambda$ is determined by field theory. This then puts bounds on \mbox{$m_h$} . These bounds can be understood as follows. The self coupling receives radiative corrections from the diagrams indicated in \begin{figure}[htb] \begin{center} \psfig{figure=self.eps,height=25mm} \end{center} \caption{Diagrams responsible for radiative corrections to self coupling. \label{figure2}} \end{figure} Fig.~\ref{figure2}. The contributions from the diagrams involving the scalar and the gauge boson loops on one hand and the fermion loops on the other, are opposite in sign. The requirement that the self coupling $\lambda$ stay positive, i.e. the vacuum remain stable under radiative corrections, puts a lower bound on \mbox{$m_h$}\ for a given value of $m_t$~\cite{ltop}. This bound however, depends on the $h t \bar t$ coupling and hence can be evaded in models with more than one Higgs doublets. \mbox{$m_h$}\ is also bounded from above by considerations of triviality. This can be understood by considering only the contributions of the scalar loops, for simplicity, to the radiative corrections to $\lambda$ . It can be shown that the self coupling then satisfies, \begin{equation} \frac{d \lambda(t)}{dt} = \frac{3}{4 \pi^2} \lambda^2(t) \label{e2} \end{equation} with $t = \ln(\Lambda /{\mathit v})$; where $\Lambda$ is the momentum scale at which the coupling $\lambda$ is evaluated. This equation needs a boundary condition to solve it, which is chosen as \begin{equation} \lambda = \lambda(\Lambda={\mathit v}) = \lambda(1) = \sqrt{2} G_F m_h^2. \label{e3} \end{equation} Then Eq.~\ref{e2} above can be solved to give \begin{equation} \lambda (t) = \frac{\lambda}{1 - {3 \lambda t}/{4 \pi^2}} \label{e4} \end{equation} This shows that $\lambda(t)$ will diverge at high scales. If we demand that the Landau pole, where $\lambda (t)$ will blow up, lies above a scale $\Lambda$, we get \begin{equation} m_h \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: \frac{893}{\sqrt{\ln (\Lambda/\mbox{$\mathit v$})}} {\mathrm GeV}. \label{e5} \end{equation} Hence the requirement that the theory be valid at large $\Lambda$ and yet be nontrivial at a scale \mbox{$\mathit v$}\ , puts an upper limit on $\lambda (\mbox{$\mathit v$})$ and hence on \mbox{$m_h$}\ due to the identification in Eq.~\ref{e1}. The above analysis, of course, has to be improved using the renormalisation group equation~\cite{sher,lindner}. As $\lambda$ becomes large, clearly perturbative methods used above must fail. In the region of large $\lambda$ the analysis has been done using lattice theory~\cite{AH}. The lower bound on \mbox{$m_h$}\ implied by the vacuum stability arguments~\cite{ltop} and the upper bound implied by the triviality considerations~\cite{sher,lindner}, depend on the value of \mbox{$m_t$}\ and the uncertainties in the nonperturbative dynamics respectively. The resulting bands, taking into account these theoretical uncertainties, are shown in \begin{figure}[htb] \begin{center} \vspace{-0.3cm} \mbox{\epsfig{file=limits.eps,height=60mm}} \vspace {- 0.7 cm} \caption{Theoretical bounds on \mbox{$m_h$}\ in the SM. \label{figure3}} \end{center} \end{figure} Fig.~\ref{figure3}~\cite{hambye}. The results can be summarised as follows: \begin{enumerate} \item If we demand that the Landau pole lies above $10^{15} (10^{18})$ GeV and there exists no new physics other than the SM upto that scale, then $\mbox{$m_h$} < 190 (130)$ GeV. \item If we assume that the SM is an effective theory only upto 1 TeV, i.e., there is some new physics at that scale then $m_h \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 800 $ GeV. As an aside let us also mention here that this upper limit of 800 GeV is of the same order as the limit obtained by requiring that $WW \rightarrow WW (ZZ \rightarrow ZZ) $ amplitude satisfies perturbative unitarity~\cite{llsmith,thacker}. \end{enumerate} \subsection {Masses and couplings in the MSSM} In the MSSM there exist two complex Higgs doublets $\Phi_1, \Phi_2$ with hypercharge $Y = \pm$ respectively. As a result there are five physical degrees of freedom left after the spontaneous symmetry breakdown. The MSSM thus contains, in all, five scalars: three neutrals out of which two are CP even states denoted by $h_0, H_0$ and one is a CP odd state denoted by A and a pair of charged Higgs bosons $H^\pm$. Thus the scalar sector of the MSSM is much richer than in the SM. $h_0$ denotes the lighter of the two CP even neutral scalars. The most general scalar potential for a two Higgs doublet model contains a large number of free parameters, essentially analogues of the self coupling $\lambda$ and $\mu^2$ term in the case of the SM. However, supersymmetry either fixes all these self couplings in terms of the gauge couplings or requires them to vanish. Hence the scalar potential for the MSSM is \begin{eqnarray} V &=& m_{11}^2 \mbox{$\Phi_1$}^\dag \mbox{$\Phi_1$} + m^2_{22} \mbox{$\Phi_2$}^\dag \phi2 - \left[ m^2_{12} \mbox{$\Phi_1$}^\dag \mbox{$\Phi_2$} + {\mathrm h.c.} \right]\nonumber\\ &&+\frac {1}{8} (g^2 + g'^2 ) \left[ (\mbox{$\Phi_1$}^\dag \mbox{$\Phi_1$})^2 + (\mbox{$\Phi_2$}^\dag \mbox{$\Phi_2$})^2 \right]\nonumber \\ &&+\frac{1}{4} (g^2 - g'^2)(\mbox{$\Phi_2$}^\dag \mbox{$\Phi_2$}) (\mbox{$\Phi_1$}^\dag \mbox{$\Phi_1$}) - \frac{1}{2} g^2 (\mbox{$\Phi_1$}^\dag \mbox{$\Phi_2$})(\mbox{$\Phi_2$}^\dag \mbox{$\Phi_1$}), \label{e6} \end{eqnarray} where $g,g'$ are the $SU(2), U(1)$ coupling constants respectively. After the SSB where the neutral members of the two doublets acquire vacuum expectation values $\mbox{$\mathit v$}_1/\sqrt{2}, \mbox{$\mathit v$}_2/\sqrt{2}$ respectively, the resulting masses of the five physical scalars that follow from the abovementioned potential, can be written in terms of two parameters which can be chosen to be $m_A, \tan \beta = \mbox{$\mathit v$}_1 / \mbox{$\mathit v$}_2$ or equivalently $m_{H^\pm}, \tan \beta$. At the tree level the five scalar masses satisfy the following inequalities: \begin{equation} \mbox{$m_{h_0}$} \leq m_Z,\;\;\; \mbox{$m_{H_0}$} > m_Z,\;\;\; m_{H^\pm} > m_W,\;\;\; \mbox{$m_{h_0}$} < \mbox{$m_{H_0}$}, m_{H^\pm} . \label{e7} \end{equation} In the decoupling limit~\cite{haber1} ($m_A \rightarrow \infty $) one finds that, independent of $\tan \beta$, all the four heavy scalars become degenerate and infinitely heavy and the mass of the lightest scalar approaches the upper bound. In this limit the couplings of the $h_0$ to matter fermions and the gauge bosons approach those of the SM higgs $h$. The interesting thing to note here is that all the masses $m_{H^\pm}, m_A, \mbox{$m_{h_0}$}$ can become large without some self coupling becoming strong, unlike the case of the SM. If we denote by $\alpha$ the mixing in the two CP even neutral fields to yield the mass eigenstates $h_0,H_0$, then the couplings of all the three neutral scalars, with the fermions and gauge bosons are given in Table~\ref{table1}. \begin{table}[htb] \begin{center} \caption{Couplings of the three neutrals with fermions and gauge bosons \label{table1}} \vspace{0.2cm} \footnotesize \begin{tabular}{|c|c|c|c|c|} \hline &&&&\\ & {$h$} & {$h_0$} &{$H_0$} & A\\ &&&&\\ \hline &&&&\\ {$b \bar b$}&{${g m_b}/{2m_W}$}&{$\sin \alpha /\cos\beta$} &{$\cos \alpha /\sin \beta$}&{$\tan \beta$}\\ &&{$\rightarrow 1$}&{$\tan \beta$}&{$\tan \beta$}\\ \hline &&&&\\ {$t \bar t$}&{${g m_t}/{2m_W}$}&{$\cos \alpha /\sin \beta$} &{$\sin \alpha /\cos \beta$}&{$\cot \beta$}\\ &&{$\rightarrow 1$}&{$\cot \beta$}&{$\cot \beta$}\\ \hline &&&&\\ {$VV$}&{$g m_V$}&{$\sin (\beta - \alpha)$} &{$\cos (\beta - \alpha)$}&0\\ &&{$\rightarrow 1$}&0&0\\ \hline \end{tabular} \end{center} \end{table} For the MSSM scalars only the additional factors relative to the SM case are written down. These couplings in the {\it decoupling limit} reduce to the ones denoted in the second line in each case. We notice that in the decoupling limit, where the upper limit on the mass of the lightest scalar is saturated, the couplings of the lightest scalar approach those for the SM higgs $h$. It should also be noted that the CP odd scalar A does not couple to a pair of gauge bosons at the tree level. The inequalities of Eq.~\ref{e7} get affected by the radiative corrections to the Higgs masses. The dominant corrections arise from loops involving top (t) and its scalar partner stop ($\tilde t$) due to the large Yukawa coupling of the top quark. There are many different methods, involving different methods of approximations~\cite{haber2}, to calculate these corrections. These depend on \mbox{$m_t$}\ as well as the masses of and the mixing between $\tilde t_L$, $\tilde t_R$ (the superpartners of $t_L$ and $t_R$). To a good approximation, the radiatively corrected upper bound on the mass of the lightest CP neutral scalar (\mbox{$m_{h_0}$}), can be written as \begin{equation} m^2_{h_0} < m_Z^2 \cos^2 2\beta + \epsilon + \epsilon_{\mathrm mix}, \label{e8} \end{equation} where \begin{eqnarray} \epsilon &=& \frac{3 g^2 m_t^4}{8 \pi^2 m^2_W} \ln \left(\frac{m^2_{\tilde t}}{m^2_t}\right) \nonumber \\ \epsilon_{\mathrm mix}&=& \frac{3 g^2 m_t^4}{8 \pi^2 m^2_W} \frac {A^2_t}{m^2_{\tilde t}} \left(1 - \frac{A^2_t}{12 m^2_{\tilde t}}\right), \label{e9} \end{eqnarray} with $A_t$ being the coefficient of the trilinear, supersymmetry breaking term, and $m_{\tilde t}$ being the common mass of the $\tilde t_L, \tilde t_R$. The second term $\epsilon_{\mathrm mix}$, even though dependent on $A_t$, can be shown to be bounded by ${9 g^2 m^4_t}/{8 \pi^2 m_W^4}$. Thus \mbox{$m_{h_0}$}\ is still bounded eventhough the bound on \mbox{$m_{h_0}$}\ of Eq.~\ref{e7} is changed by radiative corrections. Also note that the corrections will vanish in the limit of exact supersymmetry. The limits on the radiatively corrected scalar masses for the case of maximal mixing in \begin{figure}[htb] \vspace{1cm} \begin{center} \mbox{\epsfig{file=higloop.eps,height=45mm}} \vspace{-0.5cm} \end{center} \caption{Bounds on the masses of the scalars in the MSSM \label{figure4}} \end{figure} the stop sector are shown in Fig.~\ref{figure4}~\cite{abdel}. It can be seen from the results shown in the figure that the mass of the lightest scalar in MSSM is bounded by $\sim 130 $ GeV even after it is radiatively corrected. This bound does get modified in the NMSSM~\cite{kane,quiros,pandita}. Again, it has been shown that for all reasonable values of the model parameters, \mbox{$m_{h_0}$}\ is bounded by $\sim 150$ GeV. From the figure, it might seem that the current LEP bound on the mass of the \mbox{$m_{h_0}$} , \mbox{$m_{H_0}$} and $m_A$ rule out the low $\tan \beta$ values for the MSSM. However, it has been shown~\cite{dpman} that a rather minor extension of the MSSM, can help avoid this conclusion. The couplings of Table~\ref{table1} do get modified to some extent by the radiative corrections, but the general features discussed above remain unchanged. In discussing the search strategies and propspects of the MSSM scalars one has to remember the following important facts: \begin{enumerate} \item Due to the reduction of the $h_0 W W$ coupling as indicated in Table~\ref{table1}, the $h_0 \gamma \gamma$ coupling is suppressed as compared to the corresponding SM case. Of course one also has to include the contribution of the charged sparticles in the loop~\cite{abdel2}. Due to the upper limit on \mbox{$m_{h_0}$}\ the decay mode into $WW (VV)$ pair is not possible for $h_0$, due to kinematic reasons. On the other hand, for $H_0$ the suppression of the coupling to $VV$ as seen from Table~\ref{table1}, makes the decay less probable as compared to the SM case. As a result, the MSSM scalars are expected to be much narrower resonances as compared to the SM case. For example, the maximum width of $h_0$ is less than few MeV, for reasonable values of $\tan \beta$ and even for the heavier scalars $H_0, A$, the width is not more than few tens of GeV even for masses as high as 500 GeV. \item $h_0$ is much narrower than the SM Higgs. However, over a wide range of $\tan \beta, m_A$ values, the $h_0$ has dominant decay modes into Supersymmetric particles. The most interesting ones are those involving the lightest neutralinos, which will essentially give `invisible' decay modes to the $h_0, H_0$ and $A$ ~\cite{abdel1}. \item On the whole for the MSSM scalars the decay modes into fermion-antifermion pair are the dominant ones due to the point (1) above as well as the fact that the CP odd scalar A does not have any tree level couplings to $VV$. Hence, looking for the $\tau^+ \tau^-$ and $\bar b b$ final state becomes very important for the search of the MSSM scalars. \end{enumerate} It is clear from the above that the phenomenology of the MSSM scalars is much richer and more complicated than the SM case. Again, calculation of various decay widths including the higher order corrections has been done~\cite{spira,abdel2}. \section{Production and search of Higgs at Colliders} In this section we will begin by discussing the search possibilities for the SM Higgs $h$. As is clear from the discussions in the earlier section of the couplings of the Higgs, the most efficient way of producing the Higgs scalar at any Collider is through its coupling to gauge bosons or to a heavy fermion-antifermion ($t \bar t$) pair. At the hadronic colliders the following processes can contribute to the production: \begin{eqnarray} \label{e10} gg &\rightarrow & h \\ \label{e11} q \bar q' &\rightarrow & h W\\ \label{e12} q \bar q &\rightarrow& h Z \\ \label{e13} qq &\rightarrow&h q q \\ \label{e14} gg, q \bar q &\rightarrow& h t \bar t, h b \bar b. \end{eqnarray} At the Tevatron energies ( for Run-II/TeV33) the most efficient processes are those of Eqs.~\ref{e10},\ref{e11}. The potential of direct Higgs search at the Tevatron is discussed elsewhere in the proceedings~\cite{paulg}. The associated $W/Z$ can make it possible to use the dominant $b \bar b$ decay mode. However, the energy of the Tevatron is just too small to give appreciable production cross-section, except for the values of \mbox{$m_h$}\ close to its current lower limit of $92.5$ GeV~\cite{LEPCH}. Various strategies for using the $W W^*$ or the $b \bar b$ decay mode for $h$ to enhance the mass reach of the Tevatron have been suggested~\cite{tao}. At LHC energies, due to the large available gluon fluxes and the large value of \mbox{$m_t$}, Eq.~\ref{e10} is the dominant production process for all values of \mbox{$m_h$}\ upto the upper bounds discussed in section~\ref{twoone}. The total production cross-section goes from $\sim 20$ pb to $\sim 0.1$ pb as \mbox{$m_h$}\ goes from 100 to 1000 GeV. For a superheavy Higgs ($\:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\:$ 800 GeV i.e. above the highest upper bound of sec.~\ref{twoone}), the process of Eq.~\ref{e13} has significant cross-section. The higher order QCD corrections to the gluon induced higgs production are significant and have been included in the available theoretical predictions shown in \begin{figure}[htb] \begin{center} \mbox{\epsfig{file=prohiggs.eps,height=60mm}} \vspace{-0.5cm} \end{center} \caption{Production cross-section for the SM Higgs at LHC \label{figure5}} \end{figure} Fig.~\ref{figure5}~\cite{spira}. These predictions, at the LHC, have a typical theoretical uncertainty of $\sim 20\%$ due to the parton densities. Since LEP-2 has already ruled out, from direct search, $m_h < 92.5$ GeV~\cite{LEPCH}, the range of masses of interest to LHC divides neatly into two parts : $92.5 < \mbox{$m_h$} \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 140$ GeV and $\mbox{$m_h$}\ > 140$ GeV. In the first mass region the dominant decay mode of the $h$ is into a $b \bar b$ final state which has a QCD background about $10^3$ higher than the signal. Hence, in this mass range $h \rightarrow \mbox{$\gamma \gamma$}$ remains the best final state, even with a branching ratio $\sim 10^{-3}$. Even then, the resolution required for the $M_{\mbox{$\gamma \gamma$}}$ measurement has to be $\:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 1$ GeV $\simeq .1 \% \; \mbox{$m_h$}$. The new developements in the detection aspect have been detector simulations for the \mbox{$\gamma \gamma$}\ and $b \bar b$ mode for the planned detector designs. \begin{figure}[htb] \begin{center} \vspace{-0.8cm} \mbox{\epsfig{file=lowmshig.eps,height=60mm}} \end{center} \vspace{-.3cm} \caption{The expected significance level of the SM Higgs signal at LHC for the intermediate mass region\label{figure6}} \end{figure} Fig.~\ref{figure6} taken from the CMS/ATLAS technical proposal~\cite{cms1} shows the expected $S/\sqrt{B}$ for the SM Higgs in the intermediate mass range, using the $\mbox{$\gamma \gamma$} , b \bar b$ modes. The use of $b \bar b$ mode for $\mbox{$m_h$} < 100 $ GeV, is essentially achieved by using the associated production of Eq.~\ref{e11}. Once $\mbox{$m_h$}\ \:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: 140 $ GeV , the $VV^*$ or $VV$ decay mode is dominant. As a result, for $140 < \mbox{$m_h$}\ < 600$ GeV, the 4 lepton final state following from $h \rightarrow ZZ \rightarrow l^+l^-l^+l^-$ offers the so called `gold plated signal' for the Higgs. For $800 < \mbox{$m_h$} <1000$, the very forward jets produced in association with the Higgs in the process of Eq.~\ref{e13} provide a much better signature than the `gold plated signal'. However, it should be borne in mind, that the unitarity and triviality bounds of section~\ref{twoone} imply that in the SM $\mbox{$m_h$}\ \:\raisebox{-0.5ex}{$\stackrel{\textstyle<}{\sim}$}\: 600-800 $GeV. Thus we see that while the detection of an intermediate mass Higgs is difficult but feasible at LHC, it will surely require the high luminosity run. It should also be kept in mind that a low value of \mbox{$m_h$}\ seems to be preferred by the precision measurements at LEP and SLC and further that SUSY also predicts that the lightest scalar in the theory will be in this mass range. For higher \mbox{$m_h$}\ values the detection is a certainty at LHC. However, to establish such a scalar as {\it the SM} Higgs, one needs to establish \begin{enumerate} \item The scalar is CP even and has $J^P = 0^+$, \item The couplings of the scalar with the fermions and gauge bosons are proportional to their masses. \end{enumerate} This is also essential from the point of view of being able to distinguish this scalar from the lightest scalar expected in the MSSM. We see from Table~\ref{table1} that the couplings of the scalars to the fermions and gauge bosons can be quite different in the MSSM. As a matter of fact this issue has been a subject of much investigation of late~\cite{tevwg,gunion}. The Snowmass Studies~\cite{tevwg} indicate that for a light Higgs ($\mbox{$m_h$}\ = \mbox{$m_Z$}$) it is possible only to an accuracy of about 30 \%. It is in this respect that the planned \mbox{$e^+e^-$}\ colliders~\cite{physrep} can be a lot of help. At these colliders, the production processes are the same as given in Eqs.~\ref{e12}-\ref{e14} where $q (\bar q)$ are replaced by $e^- (e^+)$. This is not an appropriate place to give a complete discussion of the search prospects for the SM (and MSSM) Higgs at these colliders. But suffice it to say that if the production is kinematically allowed, detection of the Higgs at these machines is very simple as the discovery will be signalled by very striking features of the kinematic distributions. Determination of the spin of the produced particle in this case will also be simple as the expected angular distributions will be very different for even and odd parity. Even with this machine one will need a total luminosity of $200\;\; {\mathrm fb}^{-1}$, to be able to determine the ratio of $BR (h \rightarrow c \bar c) / BR(h \rightarrow b \bar b)$, to about $7 \%$~\cite{gunion}. The simplest way to determine the CP character of the scalar will be to produce $h$ in a \mbox{$\gamma \gamma$}\ collider, which are being discussed. There are also interesting invstigations~\cite{gunion} which try to device methods to determine the CP character of the scalar using hadron colliders. For MSSM Higgs the discussion of the actual search possibilities is much more involved and has been covered in other talks at this conference~\cite{expthig}. For the lightest scalar \mbox{${h_0}$}\ in the MSSM, the general discussions of the intermediate mass Higgs apply, with the proviso that the \mbox{$\gamma \gamma$}\ branching ratios are smaller for \mbox{${h_0}$}\ and hence the search that much more difficult. However, since there exist many more scalars in the spectrum now, one can cover the different regions in the parameter space by looking for $A,H$ and $H^\pm$. At low values of \mbox{$m_A$}\ these other scalars are kinematically accessible at LHC and also at the NLC. However, even after combining the information from various colliders (LEP-II, Tevatron (for the charged higgs search) and of course LHC), a certain region in the $\mbox{$m_A$} - \tan \beta$ plane remains inaccessible. This hole can be filled up only after combining the data from the CMS and ATLAS detector for 3 years of high luminosity run of LHC. Even in this case there exist large region where one will see only the single light scalar. At large \mbox{$m_A$}\ (which seem to be the values preferred by the current data on $b \rightarrow s \gamma$), the SM higgs and \mbox{${h_0}$}\ are indistinguishable as far as their couplings are concerned, as can be seen from Table~\ref{table1}. A recent study, gives the contours of constant values for the ratio $$ \frac{BR(c \bar c)/ BR(b \bar b)|_{\mbox{${h_0}$}}} {BR(c \bar c)/ BR(b \bar b)|_{h}}, $$ as well as a similar ratio for the $WW^*$ and $b \bar b$ widths as a function of $\tan \beta$ and \mbox{$m_A$} . \begin{figure}[htb] \begin{center} \vspace{-0.8cm} \mbox{\epsfig{file=test_gunion.eps,height=70mm}} \end{center} \vspace{-0.8cm} \caption{The ratios of relative branching fractions for the MSSM and SM for maximal mixing in squark sector. The specific value of the Higgs mass used is theoretically disallowed at large \mbox{$m_A$}\ and around $\tan \beta \sim 2$. \label{figure7}} \end{figure} As we can see from Fig.~\ref{figure7}~\cite{tevwg}, a measurement of this ratio to an accuracy of about $10 \%$ will allow distinction between the SM Higgs $h$ and MSSM Higgs \mbox{${h_0}$}\ for $m_A$ as large as $\sim 600 $ GeV. As stated above, NLC should therefore be able to do such a job. Certainly, the issue of being able to determine the quantum numbers and the couplings of the scalar accurately, forms the subject of a large number of investigations currently. \section{Conclusions} \begin{enumerate} \item The experiments at the upcoming hadronic colliders will be able to `discover' scalar in the entire mass range from 100 GeV to 1000 GeV, with varying degree of ease. The lower end is difficult but feasible. \item If the experiments at the Tevatron (Run-II and TeV33, should it happen) and the LHC do not find a light scalar upto $\mbox{$m_h$}\ = 130 (160) $ GeV, it will certainly rule out a class of SUSY models; viz. MSSM and models with minimal extension of the MSSM: the (N)MSSM. The only way this upper bound can be relaxed is if we enlarge the gauge group and introduce additional scales in the problem. \item If the scalar that is `found' has a mass $\:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: 800$ GeV, it is an indication that the EW symmetry breakdown happens via strongly interacting sector. \item If we {\it do} find a {\bf light} Higgs, then all we can conclude is that the SM works as an effective theory upto large scales, as discussed in section~\ref{twoone}. It will also be consistent with the indirect mass limits obtained from precision measurements. This limit is more or less insensitive to existence of SUSY (or otherwise) due to large mass scales to which sparticle masses have already been pushed by the lack of direct observation of the SUSY particles \item Since in Supersymmetric models the masses and the decay modes of the various scalars in the theory are correlated, almost all the region in the parameter space of the MSSM can be explored at the experiments at the LHC~\cite{expthig}. \item Already with the current data (particularly the information on the $b \rightarrow s \gamma$) the limits on the MSSM parameter space are such that the lightest scalar \mbox{${h_0}$}\ will be very similar in its properties to the SM Higgs $h$. Hence, it is important to devise strategies to determine the quantum numbers such as Spin, Parity, CP etc., of this observed scalar. To that end a TeV ($\:\raisebox{-0.5ex}{$\stackrel{\textstyle>}{\sim}$}\: 300 $ GeV) linear \mbox{$e^+e^-$}\ collider seems indispensible. Also, the possibilities of achieving this at the LHC/Run-II/TeV-33 need to be investigated vigorously. \end{enumerate} \section*{Acknowledgements:} It is a pleasure to thank Profs. Narasimham and Mondal for organising an excellent conference which provided the forum for many interesting discussions. This work was in part supported by the Department of Science and Technology (India) and the National Science Foundation, under NSF-grant-Int-9602567. \section*{References}
\section{ Introduction} The existence of fixed lightcone structures is one of the characteristics of classical gravitational theory. Lightcones are basically hypersurfaces which distinguish timelike separation from spacelike separation and divide spacetime into causally distinct regions. However, if gravity is to be quantized, it is natural to expect that the quantum metric fluctuations would smear out the lightcone, and the concept of a fixed lightcone structure has to be abandoned. Based upon the observation that the ultraviolet divergences of quantum field theory arise from the light cone singularities of two-point functions, and that quantum fluctuations of the spacetime metric ought to smear out the light cone, thus possibly removing these singularities, Pauli\cite{Pauli} conjectured many years ago that the ultraviolet divergences of quantum field theory might be removed if gravity is quantized. This idea was further explored by several other authors \cite{Deser,DW,ISS}. At present time, this conjecture remains unproven. If lightcones fluctuate, so do horizons, which are, of course, lightcones. The horizon fluctuations could then presumably lead to information leakage across the black hole in a way that is not allowed by classical physics. Bekenstein and Mukhanov \cite{MB} have suggested that horizon fluctuations could result in discreteness of the spectrum of black holes. Since the existence of black hole horizons is the origin of the so-called black hole information paradox, which has been widely discussed in the literature but still remains to be resolved, the study of lightcone fluctuations might help us better understand the problem. Recently the problem of light cone fluctuations has been investigated \cite{Ford95,Ford96} in a model of quantum linearized theory of gravity, where the fluctuations are produced by gravitons propagating on a background spacetime. The lightcone is smeared out if the linearized gravitational perturbations are quantized. It has been demonstrated that gravitons in a quantum state, such as a squeezed vacuum state, or a thermal state, can produce light cone fluctuations, thus smearing out the light cone. Because of the fluctuating light cone, the propagation time of a classical light pulse over distance $r$ is no longer precisely $r$, but undergoes fluctuations around a mean value of $r$. The fluctuations in the photon arrival time can also be understood as fluctuations in the velocity of light. This model has been applied to study the quantum cosmological and black hole horizon fluctuations \cite{Ford97}. It is interesting to note that recently, the quantum gravitational metric fluctuations have also been discussed within a different context, i.e., a Liouville string formulation of quantum gravity \cite{AEMN,EMN} . In this paper we shall examine light cone fluctuations in flat spacetime with nontrivial topology based upon the model proposed in Ref\cite{Ford95} . In Sec. II, we review the basic formalism and examine its gauge invariance, then derive general expressions for the vacuum graviton two-point functions in the transverse trace-free gauge. In Sec. III we study the light cone fluctuations in flat spacetimes with a compactified spatial dimension, and with a single plane boundary. Our results are summarized and discussed in Sec. VI. \section{ Basic formalism and graviton two-point function in transverse trace-free gauge} Let us consider a flat background spacetime with a linearized perturbation $h_{\mu\nu}$ propagating upon it , so the spacetime metric may be written as \begin{equation} ds^2 = g_{\mu\nu}dx^\mu dx^\nu = (\eta_{\mu\nu} +h_{\mu\nu})dx^\mu dx^\nu = dt^2 -d{\bf x}^2 + h_{\mu\nu}dx^\mu dx^\nu \, . \label{eq:metric} \end{equation} Let $\sigma(x,x')$ be one half of the squared geodesic separation for any pair of spacetime points $x$ and $x'$, and $\sigma_0(x,x')$ be the corresponding quantity in the flat background . We can expand, in the presence of the perturbation, $\sigma(x,x')$ in powers of $h_{\mu\nu}$ as \begin{equation} \sigma = \sigma_0 + \sigma_1 + \sigma_2 + \cdots \, , \label{eq:sigma} \end{equation} where $\sigma_1$ is first order in $h_{\mu\nu}$, etc. We now suppose that the linearized perturbation $h_{\mu\nu}$ is quantized, and that the quantum state $|\psi \rangle$ is a ``vacuum'' state in the sense that we can decompose $h_{\mu\nu}$ into positive and negative frequency parts $h^{+}_{\mu\nu}$ and $h^{-}_{\mu\nu}$, respectively, such that \begin{equation} h^{+}_{\mu\nu} |\psi \rangle =0, \qquad \langle \psi|h^{-}_{\mu\nu} =0 \,. \end{equation} It follows immediately that \begin{equation} \langle h_{\mu\nu} \rangle =0 \end{equation} in state $|\psi \rangle$. In general, however, $\langle (h_{\mu\nu})^2 \rangle_R \not= 0$, where the expectation value is understood to be suitably renormalized. This reflects the quantum metric fluctuations. \subsection{Basic formalism and gauge invariance} If we average the retarded Green's function, $G_{ret}(x,x') $, for a massless scalar field, over quantized metric fluctuations, we get \cite{Ford95} \begin{equation} \Bigl\langle G_{ret}(x,x') \Bigr\rangle = {{\theta(t-t')}\over {8\pi^2}} \sqrt{\pi \over {2\langle \sigma_1^2 \rangle}} \; \exp\biggl(-{{\sigma_0^2}\over {2\langle \sigma_1^2 \rangle}}\biggr)\, . \label{eq:retav} \end{equation} This form is valid for the case in which $\langle \sigma_1^2 \rangle > 0$. It reveals that the delta-function behavior of the classical Green's function, $G_{ret}$, has been smeared out into a Gaussian function peaked around the classical lightcone. This smearing can be understood as due to the fact that photons may be either slowed down or speeded up by the light cone fluctuations. Photon propagation now becomes a statistical phenomenon, with some photons traveling slower than the light on the classical spacetime, and others traveling faster. Note that the Gaussian function in Eq.~(\ref{eq:retav}) is symmetrical about the classical light cone, $ \sigma_0=0$, and so the quantum fluctuations are equally likely to produce a time advance as a time delay. Light cone fluctuations are in principle observable. It has been shown, by considering light pulses between a source and a detector separated by a distance $r$, that the mean deviation from the classical propagation time is related to $\langle \sigma_1^2 \rangle$ by \cite{Ford95} \begin{equation} \Delta t= {\sqrt{\langle \sigma_1^2 \rangle}\over r}\,. \label{eq:MDT} \ee Note, however, that $\Delta t$ is the ensemble averaged deviation, not necessarily the expected variation in flight time of two photons emitted close together in time. The latter can be much smaller than $\Delta t$ due to the fact that the gravitational field may not fluctuate significantly in the interval between the two photons. This point is discussed in detail in Ref.~\cite{Ford96}. In order to find $\Delta t$ in a particular situation, we need to calculate the quantum expectation value $\langle \sigma_1^2 \rangle$ in a chosen quantum state. For this purpose, we first have to compute $\sigma_1$ for a given classical perturbation along a certain geodesic, then average $\sigma_1^2$ over the quantized metric perturbation. If we consider a null geodesic specified by \begin{equation} dt^2=d{\bf x}^2-h_{\mu\nu}dx^{\mu}dx^{\nu}, \ee then by following the same steps as those of Ref.\cite{Ford95}, we can show that in a general gauge \begin{equation} \sigma_1 = {1\over 2}\Delta r \int_{r_0}^{r_1} h_{\mu\nu} n^{\mu} n^{\nu}\,dr\,, \end{equation} and \begin{equation} \langle \sigma_1^2 \rangle = {1\over 4}(\Delta r)^2 \int_{r_0}^{r_1} dr \int_{r_0}^{r_1} dr' \:\, n^{\mu} n^{\nu} n^{\rho} n^{\sigma} \:\, \langle h_{\mu\nu}(x) h_{\rho\sigma}(x') \rangle_R \,. \end{equation} Here $ dr=|d{\bf x}|$, $\Delta r=r_1-r_0$ and $ n^{\mu} =dx^{\nu}/dr$. The graviton two-point function, $\langle h_{\mu\nu}(x) h_{\rho\sigma}(x') \rangle_R$, is understood to be renormalized, so that it is finite when $x=x'$ and vanishes when the quantum state of the gravitons is the Minkowski vacuum state. A few comments on the derivation of Eq.~(\ref{eq:retav}) are in order here. It is obtained by averaging the Fourier representation of a $\delta$-function. It may come as a surprize that although we started with an analytic expansion of $\sigma$ in powers of $h_{\mu\nu}$, the result is not analytic as $\langle \sigma_1^2 \rangle \rightarrow 0$. This arises because we use the first order expansion of $\sigma$ in the argument of an exponential function, but afterwards retain all powers of $h_{\mu\nu}$. One can reasonably ask whether this is a valid procedure. A test of the self-consistency is to retain the $\sigma_2$ term and then follow the same procedure. The result is Eq.~(\ref{eq:retav}) with $\sigma_0^2$ replaced by $\sigma_0^2 + \sigma_2$. This has the same physical interpretation as before; the only effect of the $\sigma_2$ part is to shift the location of the mean lightcone. Thus in this order we encounter the backreaction of the gravitons in perturbing the original classical geometry to a new classical geometry. Although this is less than a complete demonstration of the validity of Eq.~(\ref{eq:retav}), it does indicate that it arises from a self-consistent calculation. In any case, the only result that we really need in the remainder of this paper is Eq.~(\ref{eq:MDT}), which may be derived either from Eq.~(\ref{eq:retav}), or else more directly by averaging the square of Eq.~(\ref{eq:sigma}). Let us now turn to the question of the gauge invariance of the formalism. Under a gauge transformation specified by \begin{equation} x^{\prime \mu}=x^{\mu}+\xi^{\mu}(x), \ee \begin{equation} h^{\prime }_{\mu\nu}(x')= h_{\mu\nu}(x)-\xi_{(\mu,\nu)}(x)\,, \ee where $\xi^{\mu}(x)$ is of order $h_{\mu\nu}$, the quantities $\sigma_1$ and $\langle \sigma_1^2 \rangle$ are not in general invariant. However, we can show that this is due to the fact that $\Delta t$ is a coordinate time interval rather than a proper time interval. To better understand the gauge invariance, let us examine a situation in which a light signal travels between two points in space labeled by $P$ and $Q$, with a classical metric perturbation $h_{\mu\nu}$ in the intervening region, as illustrated in Fig. 1. \begin{figure}[hbtp] \begin{center} \leavevmode\epsfxsize=1.6in\epsfbox{fig1.eps} \end{center} \caption{ A light ray ( dashed line ) makes a round trip travel between two points, P and Q, in space.} \label{fig=fig1} \end{figure} For simplicity, let us assume that the propagation is in the $x$-direction. We shall look at the travel time in two different gauges, or coordinate systems, primed and unprimed. For light rays traveling in $x$ direction, we have \begin{eqnarray} {dt\over dx }& =& \pm \sqrt{ 1-h_{\mu\nu}(x){dx^{\mu}\over dx}{dx^{\nu}\over dx}}\,\nonumber\\ &&\approx \pm 1 \mp{1\over2}h_{\mu\nu}(x){dx^{\mu}\over dx}{dx^{\nu}\over dx}\,. \end{eqnarray} Here the upper sign is used for outgoing light rays and the lower sign for incoming rays. So, one way travel time $\delta t$ in the unprimed gauge is \begin{eqnarray} \delta t_{P\rightarrow Q}&=&\int^{x_Q}_{x_P}\,dx\,-{1\over 2}\int^{x_Q}_{x_P}\,h_{\mu\nu}(x){dx^{\mu}\over dx} {dx^{\nu}\over dx}\,dx\nonumber\\ && =\int^{x_Q}_{x_P}\,dx\,-{1\over 2}\int^{x_Q}_{x_P}\,h^{'}_{\mu\nu}(x'){dx^{\mu}\over dx} {dx^{\nu}\over dx}\,dx -{1\over 2}\int^{x_Q}_{x_P}\,\xi_{(\mu,\nu)}(x){dx^{\mu}\over dx} {dx^{\nu}\over dx}\,dx\,, \nonumber\\ \end{eqnarray} which, within the linearized theory, can approximated as \begin{eqnarray} \delta t_{P\rightarrow Q}&=& \int^{x_Q}_{x_P}\,dx\,-{1\over 2}\int^{x^{'}_Q}_{x^{'}_P}\,h^{'}_{\mu\nu}(x'){dx^{'\mu}\over dx'} {dx^{'\nu}\over dx'}\,dx' -{1\over 2}\int^{x_Q}_{x_P}\,\xi_{(\mu,\nu)}(x){dx^{\mu}\over dx} {dx^{\nu}\over dx}\,dx, \nonumber\\ &&=\int^{x_Q}_{x_P}\,dx\,-{1\over 2}\int^{x^{'}_Q}_{x^{'}_P}\,h^{'}_{\mu\nu}(x'){dx^{'\mu}\over dx'} {dx^{'\nu}\over dx'}\,dx' -\int^{x_Q}_{x_P}\,{d\xi_x\over dx}\,dx-\int^{x_Q}_{x_P}\,{d\xi_t\over dx}\,dx \nonumber\\ &&=x_Q(t')-\xi_x(Q,t')-(\,x_P(t_0)-\xi_x(P,t_0)\,)-\xi_t(Q,t')+\xi_t(P,t_0)\nonumber\\ &&\quad -{1\over 2}\int^{x^{'}_Q}_{x^{'}_P}\,h^{'}_{\mu\nu}(x'){dx^{'\mu}\over dx'}{dx^{'\nu}\over dx'}\,dx'\,,\nonumber\\ \end{eqnarray} where we have used the fact $ dt/dx=1$ for outgoing light rays within our approximation. Similarly, we have \begin{eqnarray} \delta t_{Q\rightarrow P}&=&-\int^{x_P}_{x_Q}\,dx\,+{1\over 2}\int^{x_P}_{x_Q}\,h_{\mu\nu}(x){dx^{\mu}\over dx} {dx^{\nu}\over dx}\,dx\nonumber\\ && =\int^{x_Q}_{x_P}\,dx\,-{1\over 2}\int^{x_Q}_{x_P}\,h^{'}_{\mu\nu}(x'){dx^{\mu}\over dx} {dx^{\nu}\over dx}\,dx +{1\over 2}\int^{x_P}_{x_Q}\,\xi_{(\mu,\nu)}(x){dx^{\mu}\over dx} {dx^{\nu}\over dx}\,dx, \nonumber\\ &&=\int^{x_Q}_{x_P}\,dx\,-{1\over 2}\int^{x^{'}_Q}_{x^{'}_P}\,h^{'}_{\mu\nu}(x'){dx^{'\mu}\over dx'} {dx^{'\nu}\over dx'}\,dx' +\int^{x_P}_{x_Q}\,{d\xi_x\over dx}\,dx-\int^{x_P}_{x_Q}\,{d\xi_t\over dx}\,dx \nonumber\\ &&=x_Q(t')-\xi_x(Q,t')-(\,x_P(t_0')-\xi_x(P,t_0')\,)+\xi_t(Q,t')-\xi_t(P,t_0')\nonumber\\ &&\quad -{1\over 2}\int^{x^{'}_Q}_{x^{'}_P}\,h^{'}_{\mu\nu}(x'){dx^{'\mu}\over dx'}{dx^{'\nu}\over dx'}\,dx'\,,\nonumber\\ \end{eqnarray} using the fact that for incoming light rays, $dt/dx=-1$. Note that \begin{equation} x'_Q(t)=x_Q(t)-\xi_x(Q,t)\,, \ee \begin{equation} x'_P(t)=x_P(t)-\xi_x(P,t)\,, \ee so \begin{equation} \delta t_{P\rightarrow Q} =\delta t'_{P\rightarrow Q}-\xi_t(Q,t')+\xi_t(P,t_0)\,, \ee and \begin{equation} \delta t_{Q\rightarrow P}=\delta t'_{Q\rightarrow P}+\xi_t(Q,t')-\xi_t(P,t_0')\,. \ee It follows that the round trip travel time is \begin{equation} \Delta t=\delta t_{P\rightarrow Q}+\delta t_{Q\rightarrow P}=\Delta t' +\xi_t(P,t_0)-\xi_t(P,t_0')\,. \label{eq:corrditime} \ee Therefore, the one way travel times, $\delta t_{P\rightarrow Q}$ and $\delta t_{Q\rightarrow P} $ are, in general, not invariant unless both the source and the detector are outside the regions where gravitational perturbations $h_{\mu\nu}$ are non-zero. In that case, it is physically reasonable to set $\xi(P,t)$ and $\xi(Q,t)$ to zero. Similarly, the round trip time $\Delta t$ is invariant only if the source ( it also acts as a detector in this case ) is outside of the gravitational perturbations. However, it is interesting to note that the round trip proper time interval for the source, $\Delta \tau$, is gauge invariant. Denote the proper time intervals in two different gauges by $\Delta \tau$ and $\Delta \tau'$, and keep in mind the fact that on the world line of the source, generally, ${dx^i\over dt}<<1$. We then have \begin{eqnarray} \Delta \tau'&=&\int \sqrt{1+h'_{00}}\,dt'=\int dt'+{1\over 2}\int h'_{00} dt'\nonumber\\ &&=\Delta t'+{1\over 2}\int h_{00}dt-\int {d \xi_t\over dt}\, dt\nonumber\\ &&=\Delta t'+\xi_t(P,t_0)-\xi_t(P,t_0') +{1\over 2}\int h_{00} dt \nonumber\\ &&=\Delta t+{1\over 2}\int h_{00} dt=\Delta \tau\,, \nonumber\\ \end{eqnarray} where we have used Eq.~(\ref{eq:corrditime}). This shows that we should really consider how proper time rather than the coordinate time is affected by light cone fluctuations. However, the calculation of the proper time in a general gauge is a rather difficult task, because the source (and detector) may not be at rest with respect to the chosen coordinate system, and thus in general the emission and the subsequent reception may not happen at the same point in space. To find the proper time, we have to integrate along the geodesic between two events, the emission and the subsequent reception. In general, there is a Doppler shift due to fluctuations in the positions of the source and the mirror. However, the analysis can be greatly simplified if we adopt the transverse-tracefree ( TT ) gauge, which is specified by the conditions \begin{equation} h^j_j = \partial_j h^{ij} = h^{0\nu} = 0\, . \label{eq:the TT} \end{equation} To see this, let us examine the geodesic equations for a test particle \begin{equation} {d^2x^{\mu}\over d^2\lambda}=\,-\Gamma^{\mu}_{\rho\sigma}\,{dx^{\rho}\over d\lambda} {dx^{\rho}\over d\lambda}\,, \ee which, when written in term of derivatives with respect to coordinate time $t$, becomes \begin{equation} {d^2x^{\mu}\over d^2 t}\,+\Gamma^{\mu}_{\rho\sigma}\,{dx^{\rho}\over dt}{dx^{\sigma}\over dt}\, -\Gamma^{t}_{\rho\sigma}\,{dx^{\mu}\over dt}\,{dx^{\rho}\over dt}{dx^{\sigma}\over dt}=0\,. \ee For a non-relativistic test particle, ${dx^i\over dt}<<1$, so, to the leading order, \begin{equation} {d^2x^i\over d^2 t}\,\approx \Gamma^{i}_{tt}\,. \ee But, in the TT gauge, $\Gamma^{i}_{tt}=0$. Therefore, from the above equation, we can see that if the test particle is at rest at $t=0$, then it will subsequently always remain at rest \cite{MTW}. So, if we are considering the emission and reflection of a light signal between two points (particles) in the TT gauge, then the proper time $\delta \tau$ between emission and reception (after reflection) of the signal is related with the coordinate time by \begin{equation} \delta \tau=\int \sqrt{g_{tt}}dt=\int \sqrt{(1+h_{00})}dt=\int dt=\delta t\,. \ee Here we have appealed to the fact that $h_{00}=0$ in the TT gauge. Therefore, the coordinate time for the round trip in the TT gauge is the proper time, and $\Delta t $ calculated from Eq.~(\ref{eq:MDT}) in the TT guage is actually a gauge invariant quantity. In this gauge, the mean squared fluctuation in the geodesic interval function reduces to \begin{eqnarray} \langle \sigma_1^2 \rangle &&= {1\over 4}(\Delta r)^2 \int_{r_0}^{r_1} dr \int_{r_0}^{r_1} dr' \:\, n^i n^j n^k n^m \:\, \langle h_{ij}(x) h_{km}(x') \rangle_R \, \nonumber\\ &&= {1\over 8}(\Delta r)^2 \int_{r_0}^{r_1} dr \int_{r_0}^{r_1} dr' \:\, n^i n^j n^k n^m \:\, \langle h_{ij}(x) h_{km}(x')+ h_{ij}(x') h_{km}(x) \rangle_R \,. \label{eq: interval} \end{eqnarray} Here $ n^i =dx^i/dr$ is the unit three-vector defining the spatial direction of the geodesic. \subsection{Graviton two-point function in transverse trace-free gauge} If we work in the TT gauge, the gravitational perturbations have only spatial components $h_{ij}$ and they may be quantized using a plane wave expansion as \begin{equation} h_{ij} = \sum_{{\bf k},\lambda}\, [a_{{\bf k}, \lambda} e_{ij} ({{\bf k}, \lambda}) f_{\bf k} + H.c. ]. \end{equation} Here H.c. denotes the Hermitian conjugate, $\lambda$ labels the polarization states, and \begin{equation} f_{\bf k} = (2\omega (2\pi)^3)^{-{1\over 2}} e^{i({\bf k \cdot x} -\omega t)} \ee is the mode function, where \begin{equation} \omega=|{\bf k}|, \qquad |{\bf k}|=(k_x^2+k_y^2+k_z^2)^{{1\over2}}, \ee and the $e_{\mu\nu} ({{\bf k}, \lambda})$ are polarization tensors. ( Units in which $32\pi G =1$, where $G$ is Newton's constant and in which $\hbar =c =1$ will be used in this paper.) Now we shall first calculate the Minkowski spacetime Hadamard function for gravitons in the transverse tracefree gauge. Let us define \begin{equation} G^{(1)}_{ijkl}(x,x')=\langle 0| h_{ij}(x) h_{kl}(x')+ h_{ij}(x') h_{kl}(x)|0 \rangle \,. \end{equation} Then we have \begin{equation} G^{(1)}_{ijkl}(x,x')=\frac{2 Re}{(2\pi)^3}\int\,d^3{\bf k}\sum_{\lambda} \, e_{ij} ({{\bf k}, \lambda}) e_{kl} ({{\bf k}, \lambda}) {1\over{2 \omega}}e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega(t-t')}\,. \end{equation} Equation~(\ref{eq:polsum}) in the Appendix for the summation of polarization tensors in the transverse tracefree gauge gives \begin{eqnarray} \sum_{\lambda}\, e_{ij} ({{\bf k}, \lambda}) e_{kl} ({{\bf k}, \lambda})&=&\delta_{ik}\delta_{jl} +\delta_{il}\delta_{jk}-\delta_{ij}\delta_{kl} +\hat k_i\hat k_j \hat k_k\hat k_l+\hat k_i \hat k_j \delta_{kl} \nonumber\\ &&+\hat k_k \hat k_l \delta_{ij}-\hat k_i \hat k_l \delta_{jk} -\hat k_i \hat k_k \delta_{jl}-\hat k_j \hat k_l \delta_{ik}-\hat k_j \hat k_k \delta_{il}\,, \end{eqnarray} where \begin{equation} \hat k_i=\frac{ k_i}{ k}\,. \end{equation} We find that $ G^{(1)}_{ijkl}(x,x')$ can be expressed as \cite{Footnote} \begin{eqnarray} G^{(1)}_{ijkl}(x,x')&&=2 Re \,(\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk}-\delta_{ij}\delta_{kl} + D_{ij} ) \times {1\over{(2\pi)^3}}\int\, d^3{\bf k}{1\over{2 \omega}} e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega(t-t')}\nonumber\\ &&=2 Re \,(\delta_{ik}\delta_{jl}+ \delta_{il}\delta_{jk}-\delta_{ij}\delta_{kl} +D_{ij})\times \langle 0|\phi(x)\phi(x') |0 \rangle\,,\ \end{eqnarray} where we have defined a formal operator \begin{equation} D_{ij}=\left( {\partial_i\partial^{\prime}_j\over{\nabla^2}}\delta_{kl}+ {\partial_k\partial^{\prime}_l\over{\nabla^2}}\delta_{ij} - {\partial_i\partial^{\prime}_k\over{\nabla^2}}\delta_{jl} - {\partial_i\partial^{\prime}_l\over{\nabla^2}}\delta_{jk} - {\partial_j\partial^{\prime}_l\over{\nabla^2}}\delta_{ik} -{\partial_j\partial^{\prime}_k\over{\nabla^2}}\delta_{il} +{\partial_i\partial_j^{\prime}\partial_k\partial_l^{\prime}\over\nabla^4} \right), \end{equation} and $ \langle 0|\phi(x)\phi(x') |0 \rangle$ is the usual scalar field two-point function. Here the formal operator $\nabla^{-2}$ should be understood in the sense of a Green's function, but when we do our calculations in momentum space its effect is to bring in a factor of $k^{-2}$. The combination of these results with Eq.~(\ref{eq: interval} ) gives \begin{equation} \langle \sigma_1^2 \rangle = { 1\over 4}(\Delta r)^2 \int_{r_0}^{r_1} dr \int_{r_0}^{r_1} dr' \:\,\left( 1-{ 2 ({\bf \nabla }\cdot {\bf n})({\bf \nabla}^{\prime} \cdot {\bf n})\over{\nabla^2} } +{ ({\bf \nabla }\cdot {\bf n})^2({\bf \nabla}^{\prime} \cdot {\bf n})^2\over{\nabla^4} } \right) \langle \phi(x)\phi(x') \rangle_R\,.\nonumber\\ \end{equation} Introduce two functions $F_{ij}(x,x')$ and $H_{ijkl}(x,x')$ by \begin{eqnarray} F_{ij}(x,x')&&= Re {\partial_i\partial^{\prime}_j\over{\nabla^2}} \langle 0|\phi(x)\phi(x') |0 \rangle \nonumber\\ &&= Re {\partial_i\partial^{\prime}_j\over{\nabla^2}} {1\over{(2\pi)^3}}\int\, d^3{\bf k}{1\over{2 \omega}}e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega(t-t')}\nonumber\\ &&={Re\over{(2\pi)^3}}\int\, d^3{\bf k}{k_ik_j\over{2 \omega^3}}e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega(t-t')} \,, \end{eqnarray} and \begin{eqnarray} H_{ijkl}(x,x')&&= Re {\partial_i\partial^{\prime}_j\partial_k\partial_{l}^{\prime} \over{\nabla^4}} \langle 0|\phi(x)\phi(x') |0 \rangle \nonumber\\ &&= Re {\partial_i\partial^{\prime}_j\partial_k\partial_l^{\prime}\over{\nabla^4}} {1\over{(2\pi)^3}}\int\, d^3{\bf k}{1\over{2 \omega}}e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega(t-t')}\nonumber\\ &&={Re\over{(2\pi)^3}}\int\, d^3{\bf k}{k_i k_j k_k k_l\over{2 \omega^5}}e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega(t-t')} \,. \end{eqnarray} $G^{(1)}_{ijkl}$ can be expressed as \begin{eqnarray} G^{(1)}_{ijkl}&=& 2F_{ij}\delta_{kl} +2F_{kl}\delta_{ij}-2F_{ik}\delta_{jl}-2F_{il}\delta_{jk}-2F_{jl}\delta_{ik} -2F_{jk}\delta_{il}+2H_{ijkl}\nonumber\\ &&\quad +2D^{(1)}(x,x')(\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk}-\delta_{ij}\delta_{kl}), \end{eqnarray} where \begin{equation} D^{(1)}(x,x')= -{1\over{8\pi^2\sigma_0^2}} \ee is the usual Hadamard function for massless scalar fields with $2\sigma_0^2=(t-t')^2-({\bf x-x' })^2$, and $F_{ij}(x,x')$ and $H_{ijkl}(x,x')$, which will be calculated in the Appendix, are given by \begin{equation} F_{ij}(x,x')=-{1\over{(2\pi)^2}}\partial_i\partial_j'\,\left[{1\over 2}\ln ( R^2-\Delta t^2 ) +{\Delta t \over 4R}\ln \Bigg( \frac{R+\Delta t}{R-\Delta t}\Bigg)^2 \right]\,, \label{eq:Ffunc} \ee and \begin{eqnarray} H_{ijkl}(x,x')&=&{1\over 96\pi^2}\partial_i\partial_j'\partial_k\partial_l'\, \Bigg[ (R^2+3\Delta t^2)\ln (R^2-\Delta t^2)^2\nonumber\\ \quad && + \left(3R\Delta t+{\Delta t^3\over R}\right)\ln \Bigg( \frac{R+\Delta t}{R-\Delta t}\Bigg)^2\Bigg]\nonumber\,.\\ \label{eq:Hfunc} \end{eqnarray} Here $R=|{\bf x}-{\bf x}'|$. \section{Lightcone fluctuations in flat spacetime with nontrivial topologies or boundaries} In this section, we study lightcone fluctuations in two cases: flat spacetime with a compactified spatial section, and with a single plane boundary. \subsection{ Flat spacetime with a compactified spatial section} Let us now assume that the spacetime is flat but compactified in the $z$ direction with a periodicity length $L$ ( ``circumference of the universe'' ). This means the spatial points $z$ and $z+L$ are identified. The effect of the space closure is to restrict the field modes to a discrete set \begin{equation} f_{\bf k} = (2\omega (2\pi)^2L)^{-{1\over 2}} e^{i({\bf k \cdot x} -\omega t)} \label{eq:mode1} \ee with \begin{equation} k_z={2\pi n\over L}, \qquad n=0,\pm 1, \pm 2, \pm 3,... \ee We now analyze the lightcone fluctuations, assuming that the gravitons are in the new vacuum state $|0_L\rangle$ associated the discrete modes of Eq.~(\ref{eq:mode1} ). First consider a light ray along $z$ direction, i.e. along the direction of compactification (Fig.2 ), propagating from point $(0,0,a)$ to point $(0,0,b)$ in space, then, we have from Eq.~(\ref{eq: interval}) \begin{eqnarray} \langle \sigma_1^2 \rangle &&= {1\over 8}(b-a)^2 \int_{a}^{b} dz \int_{a}^{b} dz' \:\, \langle 0_L| h_{zz}(x) h_{zz}(x')+ h_{zz}(x') h_{zz}(x)|0_L \rangle_R \,\nonumber\\ &&= {1\over 8}(b-a)^2 \int_{a}^{b} dz \int_{a}^{b} dz'\, G_{zzzz}^{(1) R}(t,0,0,z,\,t',0,0,z'). \label{eq: interval1} \end{eqnarray} Here we have defined \begin{eqnarray} &&G^{(1)R}_{zzzz}(x,x')=\langle 0_L| h_{zz}(x) h_{zz}(x')+ h_{zz}(x') h_{zz}(x)|0_L \rangle_R\nonumber\\ &&\qquad=\langle 0_L| h_{zz}(x) h_{zz}(x')+ h_{zz}(x') h_{zz}(x)|0_L \rangle- \langle 0| h_{zz}(x) h_{zz}(x')+ h_{zz}(x') h_{zz}(x)|0 \rangle\,, \end{eqnarray} and the integral is to be carried out along the geodesic. \vskip.25in \begin{figure}[hbtp] \begin{center} \leavevmode\epsfxsize=1.8in\epsfbox{fig2.eps} \end{center} \caption{ A light ray ( dashed line ) propagates in the direction of compactification in a cylindrical ``universe'' from point $(t,0,0,a)$ to point $ (t',0,0,b)$. Here only two spatial dimensions are plotted. } \label{fig=fig2} \end{figure} If we adopt the notation \begin{equation} (t,0,0,z,\,t',0,0,z')\equiv(t,z,\,t',z')\,, \ee the renormalized two-point function can be found by using the method of images to be \begin{eqnarray} G_{zzzz}^{(1) R}(t,z,\,t',z')&&={\sum_{n=-\infty}^{+\infty}}^{\prime}G_{zzzz}^{(1) }(t,z,\,t',z'+nL)\nonumber\\ &&=2\,{\sum_{n=-\infty}^{+\infty}}^{\prime}\,\Biggl(D^{(1)}(t,z,\, t', z'+nL)-2F_{zz}(t,z,\,t',z'+nL) \nonumber\\ &&\quad +H_{zzzz}(t,z,\,t',z'+nL)\Biggr)\,,\nonumber\\ \end{eqnarray} where the prime on the summation indicates that the $n=0$ term is omitted. Substituting $R_t=0$ into Eq.~(\ref{eq:tpFunc}) in the Appendix and replacing $\Delta x$ by $\Delta z$, we have \begin{eqnarray} G_{zzzz}^{(1) R}(t,z,\,t',z')&&=-{2\over \pi^2}{\sum_{n=-\infty}^{+\infty}}^{\prime} \Bigg[ \frac{\Delta t^2}{(\Delta z-nL)^4} +\frac{\Delta t^3}{4(\Delta z-nL)^5}\ln\left( \frac{\Delta z-nL-\Delta t}{\Delta z-nL+\Delta t} \right)^2 \nonumber\\ && - {2\over 3(\Delta z-nL)^2}-\frac{\Delta t}{4(\Delta z-nL)^3}\ln\left( \frac{\Delta z-nL-\Delta t}{\Delta z-nL+\Delta t} \right)^2 \Bigg]. \label{eq:TPF1} \end{eqnarray} For the null geodesic \begin{equation} \Delta t=\Delta z, \ee we get, after an evaluation of the integral, \begin{eqnarray} &&\int_{a}^{b} dz \int_{a}^{b} dz'\, G_{zzzz}^{(1) R}(t,z,\,t',z')|_{\Delta t=\Delta z }\nonumber\\ &&\quad ={1\over12\pi^2 }{\sum_{n=-\infty}^{+\infty}}^{\prime} \Bigg[{8\epsilon^2 (n^2-2\epsilon^2)\over {(n^2-\epsilon^2)^2}} +\frac{(n+2\epsilon)^3}{2(n+\epsilon)^3}\ln\left(1+{2\epsilon\over n}\right)^2 + \frac{(n-2\epsilon)^3}{2(n-\epsilon)^2}\ln\left(1-{2\epsilon\over n}\right)^2 \Bigg]\nonumber\\ &&\quad ={1\over12\pi^2 }{\sum_{n=1}^{+\infty}} \Bigg[{16\epsilon^2 (n^2-2\epsilon^2)\over {(n^2-\epsilon^2)^2}} +\frac{(n+2\epsilon)^3}{(n+\epsilon)^3}\ln\left(1+{2\epsilon\over n}\right)^2 + \frac{(n-2\epsilon)^3}{(n-\epsilon)^3}\ln\left(1-{2\epsilon\over n}\right)^2 \Bigg]\nonumber\\ &&\quad\equiv{1\over12\pi^2 }{\sum_{n=1}^{+\infty}}f(n,\epsilon)\nonumber\,,\\ \label{eq:series1} \end{eqnarray} where we have defined \begin{equation} \epsilon\equiv{(b-a)\over L}={r\over L}\,, \ee and \begin{equation} f(n,\epsilon) \equiv {16\epsilon^2 (n^2-2\epsilon^2)\over {(n^2-\epsilon^2)^2}}+ \frac{(n+2\epsilon)^3}{(n+\epsilon)^3}\ln\left(1+{2\epsilon\over n}\right)^2 + \frac{(n-2\epsilon)^3}{(n-\epsilon)^3}\ln\left(1-{2\epsilon\over n}\right)^2\,. \ee It appears that there is a singularity in the summand $f(n,\epsilon)$ whenever $n=\epsilon$, i.e., whenever the distance $r$ is an integer multiple of $L$. However this singularity is illusionary, as it should be from a physical point of view since there is nothing special when $n=\epsilon$. This can be seen if we expand the summand at the point $\epsilon=n$ to get \begin{equation} f(n,\epsilon)\approx {19\over 3}+{27\over 4}\ln(3)+\frac{27\ln(3)+68}{8n}(\epsilon-n)+O((\epsilon-n)^2)\,. \ee So, $f(n,\epsilon)$ is finite as $\epsilon$ approaches $n$. Note also that $2\epsilon=n$ is also not a singularity. The summation converges, as the asymptotic form of $f(n.\epsilon)$ as $n\rightarrow \infty$ is \begin{equation} f(n,\epsilon)\sim {32\epsilon^2\over n^2} + O(n^{-4})\,. \ee However, a generic closed form result for the summation is hard to find. So we now discuss two special cases. The first is the one in which the distance traversed by the light ray is much less than the periodicity length, $ b-a \ll L$. Then we get \begin{equation} \int_{a}^{b} dz \int_{a}^{b} dz'\, G_{zzzz}^{(1) R}(x,x') \approx\sum_{n=1}^{+\infty}{8\epsilon^2\over3 \pi^2}{1\over n^2} ={4\epsilon^2\over9}\,. \ee Substitution of this result into Eq.~(\ref{eq: interval1}) yields \begin{equation} \langle \sigma_1^2 \rangle \approx {r^4\over 18L^2 }. \ee Therefore the mean deviation from the classical propagation time is \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r} \approx {1\over 3\sqrt{2} } \,{r\over L}\,. \label{eq:t2} \ee Since we are working in Natural Units, this result reveals that the mean deviation in travel time is less than the Planck time and grows linearly with increasing $r$ when $r$ is small compared to the periodicity length $L$ of the universe. If $\epsilon\gg 1$, i.e., $r\gg L$, the light loops around the `` universe '', and summation Eq~(\ref{eq:series1}) can be approximated by the following integral \begin{eqnarray} \int_{a}^{b} dz \int_{a}^{b} dz'\, G_{zzzz}^{(1) R}(t,z,\,t',z')&&\approx {\epsilon\over 12\pi^2}\int_{1/\epsilon}^{\infty}\,dx \Bigg[\frac{(x+2)^3}{(x+1)^3}\ln\left(1+{2\over x}\right)^2 + \frac{(x-2)^3}{(x-1)^3}\ln\left(1-{2\over x}\right)^2 \nonumber\\ && \qquad \qquad \qquad \quad +{16(x^2-2)\over {(x^2-1)^2}}\Bigg]\,. \end{eqnarray} Evaluating the integral with the aid of the computer algebra package Maple, series expanding the result and keeping the leading terms only, we arrive at \begin{equation} \int_{a}^{b} dz \int_{a}^{b} dz'\, G_{zzzz}^{(1) R}(t,z,\,t',z') \approx \epsilon-{8\ln(2\epsilon)\over 3\pi^2 }\,. \ee Therefore the mean deviation from the classical propagation time is \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r} \approx {1\over 2\sqrt{2}} \,\sqrt{{r\over L}}, \ee where $r$ is assumed to be much greater than $L$. So the lightcone fluctuations can, in principle, get as large as one would like if the light ray travels around and around. This is interesting in the sense that it suggests that a fluctuation which is much greater than the Planck scale could be achieved. Now we turn to the case where the light ray moves along the direction perpendicular to that of compactification, for instance, along $x$ direction. If the light ray travels from point $(a,0,0)$ to point $(b,0,0)$, as illustrated in Fig. 3, then \begin{eqnarray} \langle \sigma_1^2 \rangle &&= {1\over 8}(b-a)^2 \int_{a}^{b} dx \int_{a}^{b} dx' \:\, \langle 0_L| h_{xx}(x) h_{xx}(x')+ h_{xx}(x') h_{xx}(x)|0_L \rangle_R \,\nonumber\\ &&={1\over 8}(b-a)^2 \int_{a}^{b} dx \int_{a}^{b} dx'\, G_{xxxx}^{(1)R }(t,x,0,0,\,t',x',0,0), \,\nonumber\\ &&= {1\over 8}(b-a)^2 \int_{a}^{b} dx \int_{a}^{b} dx'\,{\sum_{n=-\infty}^{+\infty}}^{\prime} G_{xxxx}^{(1) }(t,x,0,0,\,t',x',0,nL)\,. \end{eqnarray} \vskip.15in \begin{figure} \begin{center} \leavevmode\epsfxsize=1.8in\epsfbox{fig3.eps} \end{center} \caption{ A light ray ( dashed line ) propagates perpendicular to the direction of compactification in a cylindrical ``universe'' from point $(t,a,0,0)$ to point $ (t',b,0,0)$. Here only two spatial dimensions are plotted. } \label{fig=fig3} \end{figure} Let us now define \begin{equation} \rho=x-x',\quad\quad b-a=r, \ee then if we use Eq.~(\ref{eq:tpFunc}) in the Appendix and bear in mind the fact that for the light ray $\Delta t=\Delta x$, we have \begin{equation} G_{xxxx}^{(1)R }(t,x,0,0,\,t',x',0,nL)\equiv g_1(\rho)+g_2(\rho)\,, \label{eq:G1} \ee where \begin{equation} g_1 = 2 {\sum_{n=-\infty}^{+\infty}}^{\prime}- {\displaystyle \frac {1}{8\pi^2}} \,{\displaystyle \frac {\rho^{2}\,(nL)^{4}}{(\rho^{2} + (nL)^{2})^{4}}} - {1\over3\pi^2 } {\displaystyle \frac {\rho^{6}}{(\rho^{2} + (nL)^{2 })^{4}}} + {\displaystyle \frac {47}{12\pi^2}} \,{\displaystyle \frac {\rho^{4}\,(nL)^{2}}{(\rho^{2} + (nL)^{2})^{4}}} \,, \ee and \begin{eqnarray} g_2=-2{\sum_{n=-\infty}^{+\infty}}^{\prime} \frac {1}{2\pi^2} \,\ln\left( \frac {\sqrt{\rho^{2} + (nL)^2} + \rho}{\sqrt{\rho^{2} + (nL)^{2}} - \rho}\right )^2 &&\Biggl[ - \frac {1}{16} \, \frac { \rho\,(nL)^6 }{ (\rho^{2} + (nL)^{2})^{(9/2)} } - \frac {3}{4} \, \frac { \rho^3\,(nL)^4 }{ (\rho^{2} + (nL)^{2})^{(9/2)} } \nonumber\\ && \quad + \frac {3}{2} \,\frac { \rho^{5}\,(nL)^2 } { (\rho^2 + (nL)^2 )^{(9/2)} }\Biggr]\,. \nonumber\\ \label{eq:G3} \end{eqnarray} We can clearly see that $G_{xxxx}^{(1)R}$ is an even function of $\rho$, so, \begin{equation} \int_{a}^{b} dx \int_{a}^{b} dx'\, G_{xxxx}^{(1)R }(t,x,0,0,\,t',x',0,0)= 2\int_0^r d\rho (r-\rho)(g_1+g_2) \,. \ee Performing the integration (integrate by parts for those terms involving logarithmic function), we arrive at \begin{eqnarray} &&2\int_0^r d\rho (r-\rho)(g_1+g_2)\nonumber\\ &&\quad ={2\over\pi^2 }{\sum_{n=-\infty}^{+\infty}}^{\prime}\Bigg[ -\frac{\epsilon^4}{2(\epsilon^2+n^2)^2}-\frac{\epsilon^2n^2}{4(\epsilon^2+n^2)^2} \nonumber\\ &&\quad\quad \quad\quad+\frac{8\epsilon^5+ 8n^2\epsilon^3+3n^4\epsilon}{24(n^2+\epsilon^2)^{5/2}} \ln\left({\sqrt{n^2+\epsilon^2}+ \epsilon}\over{\sqrt{n^2+\epsilon^2}- \epsilon} \right)\Bigg]\,,\nonumber\\ \label{eq: series2} \end{eqnarray} where $\epsilon=r/L$ as before. The above series can be shown to be convergent. Yet a result in closed form is not easy to find. Let us first examine the case in which $ r\ll L$, where \begin{equation} \int_{a}^{b} dx \int_{a}^{b} dx'\, G_{xxxx}^{(1) R}(x,x')\approx \sum_{n=1}^{+\infty}{64\epsilon^6\over 45\pi^2}{1\over n^6} ={64\pi^4\epsilon^6\over 45^2\times21}\,. \ee Here we have used \begin{equation} \sum_{n=1}^{+\infty}{1\over n^6}={\pi^6\over 45\times 21}\\. \ee Thus the mean deviation from the classical propagation time is \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r}\approx \sqrt{{{2\over21 }}} {2\pi^2\over 45}\, \left({r\over L}\right)^3\,. \ee This result holds in the small $\epsilon$ regime. The time deviation is much smaller than that for light rays propagating along the compactification direction ( compare with Eq.~(\ref{eq:t2}) ). This reveals that light cone fluctuations due to topology change are more likely to be felt in the direction of compactification than in the transverse direction, if we perform local experiments in which $r$, the distance between the source and the detector, is very small as compared to $L$, the periodicity length. We now turn our attention to the case in which $r\gg L$, i.e., $\epsilon\gg 1$. Here it is easy to see that the summation in Eq.~(\ref{eq: series2}) can be approximated by the following integral \begin{eqnarray} &&\int_{a}^{b} dx \int_{a}^{b} dx'\, G_{xxxx}^{(1) R}(t,x,0,0,\,t',x',0,0)\nonumber\\ && \quad \quad \approx {4\epsilon\over\pi^2 }\int_{1/\epsilon}^\infty d x \Bigg[-\frac{1}{2(1+x^2)^2}-\frac{ x^2}{4(1+x^2)^2}\nonumber\\ &&\quad \quad\quad\quad\quad\quad\quad\quad +\frac{8+ 8x^2+3x^4}{24(x^2+1)^{5/2}} \ln\left({\sqrt{x^2+1}+ 1}\over{\sqrt{x^2+1}- 1} \right)\Bigg]\,.\nonumber\\ \, \end{eqnarray} If we perform the integral and series expand the result, we have, to the order of $O(\epsilon)$, \begin{eqnarray} &&\int_{a}^{b}dx \int_{a}^b dx'\, G_{xxxx}^{(1) R}(t,x,0,0,\,t',x',0,0) \nonumber\\ && \quad \quad \approx c_1^2\epsilon -c_2^2\ln(\epsilon)\,,\nonumber\\ \end{eqnarray} where $c_1$ and $c_2$ are constants given, respectively, by \begin{equation} c_1^2=\int_0^\infty\,dx \frac{\ln(x+\sqrt{x^2+1})}{x\sqrt{x^2+1}}\approx 2.468\,. \ee and \begin{equation} c_2^2={8\over 3\pi^2}\,. \ee Therefore we have for the mean squared geodesic interval fluctuation \begin{equation} \langle \sigma_1^2 \rangle \approx {1\over 8}\left({c_1r\over \pi }\right)^2\epsilon\,, \ee and the mean deviation from the classical propagation time is \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r}\approx {c_1\over \pi}\,{1\over 2\sqrt{2}} \sqrt{{r\over L }}. \ee This result applies in the regime where $r\gg L$. Here we have the same functional dependence on $r$ as in the case where light rays loop around the compactified dimension many times. The only difference lies in the proportionality constants. In fact, here the mean time deviation is also smaller than that for light rays traveling in the direction of compactification, since the numerical constant $ c_1/\pi\approx 0.5$. \subsection{Single plane boundary} Let us assume that there is a single plane boundary located at $z=0$ in space such that metric perturbations satisfy the following Neumann boundary condition (The reason that we use the Neumann boundary condition instead of the Dirichlet boundary condition here is to get a positive $\langle \sigma_1^2 \rangle$. ) \begin{equation} \partial_z h_{jk}|_{z=0}=0\,. \ee In the presence of the boundary, the field mode no longer has the form of Eq.~(\ref{eq:mode1} ) but becomes \begin{equation} f_{\bf k} = (\omega (2\pi)^2\pi)^{-{1\over 2}} e^{i({\bf k_t \cdot x_t} -\omega t)}\cos(k_zz), \label{eq: mode2} \ee where ${\bf k_t}$ and ${\bf x_t}$ denote the components of ${\bf k} $ and ${\bf x}$, respectively, in directions parallel to the boundary. Now if we assume that the gravitons are in the vacuum state $|0'\rangle $ associated with the modes of Eq.~(\ref{eq: mode2}), we have, for a light ray propagating perpendicular to the boundary from point $(0,0,a)$ to $(0,0,b)$, \begin{eqnarray} \langle \sigma_1^2 \rangle &&= {1\over 8}(b-a)^2 \int_{a}^{b} dz \int_{a}^{b} dz' \:\, \langle 0'| h_{zz}(x) h_{zz}(x')+ h_{zz}(x') h_{zz}(x)|0' \rangle_R \,\nonumber\\ &&=\int_{a}^{b} dz \int_{a}^{b} dz' \:\, G_{zzzz}^{(1)R}(t,0,0,z,\,t',0,0,z') \,. \label{eq: interval2} \end{eqnarray} \vskip.25in \begin{figure} \begin{center} \leavevmode\epsfxsize=1.8in\epsfbox{fig4.eps} \end{center} \caption{ A light ray ( dashed line )propagates in the direction perpendicular to the plane boundary, starting $a$ distance away from the boundary} \label{fig=fig4} \end{figure} Here the renormalized graviton two point function $ G_{zzzz}^{(1)R}(x,x') $ can be found by the method of images as usual and the only difference is an overall sign change as we go from the Dirichlet boundary condition to the Neumann boundary condition. The reason for this is that, to satisfy the Neumann boundary condition, we need to add the image term instead of subtracting it as in the case of the Dirichlet boundary condition. So, $ G_{zzzz}^{(1)R}(x,x') $ may be obtained by picking out the $n=0$ term in Eq.~(\ref{eq:TPF1}) and setting $\Delta z=z+z'$ to get \begin{eqnarray} G_{zzzz}^{(1)R}(t,0,0,z,\,t',0,0,z')&=&-\frac{2(t-t')^2}{\pi^2(z+z')^4} -\frac{(t-t')^3}{2\pi^2(z+z')^5}\ln\left( \frac{z+z'-(t-t')}{z+z+(t+t')} \right)^2 \nonumber\\ && + {4\over3\pi^2 (z+z')^2}+\frac{(t-t')}{2\pi^2(z+z')^3}\ln\left( \frac{z+z'-(t-t')}{z+z'+(t+t')} \right)^2\,. \end{eqnarray} Substituting this result into Eq.~(\ref{eq: interval2}) and performing the integration, we finally get \begin{equation} \langle \sigma_1^2 \rangle =\frac{(b-a)^3\left[ b^2-a^2+(a^2+4ab+b^2)\ln({b\over a}) \right]} {24\pi^2(b+a)^3}\,. \ee Note that this result is always greater than zero. However, had we chosen the Dirichlet boundary condition, we would have that $\langle \sigma_1^2 \rangle < 0$. Recall that the formalism which we are using applies only if $\langle \sigma_1^2\rangle > 0$. When the light ray starts very close to the boundary such that $a\ll r$, we have \begin{equation} \langle \sigma_1^2 \rangle\approx {r^4\over 24\pi^2}\left(1+\ln(r/a)\right)\,. \ee The mean deviation in travel time is \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r}=\sqrt{{ 1+\ln(r/a)\over 24\pi^2}}\,, \label{eq:t1} \ee which diverges as $a$ approaches 0. This is not surprising since the energy density of a quantized field blows up on the boundary. However, it has been shown recently \cite{Ford98} that, if one treats the boundaries as quantum objects with a nonzero position uncertainty, the singularity in energy density is removed. The result, Eq.~(\ref{eq:t1}), applies whenever $r\gg a$. The other limit is when $ r\ll a$, where the mean squared fluctuation in the geodesic interval function is approximated as \begin{equation} \langle \sigma_1^2 \rangle\approx {r^4\over 24\pi^2 a^2}\,, \ee consequently, the mean deviation in time is given by \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r}={1\over 2\sqrt{6}\pi}{r\over a}\,. \ee We now consider a null geodesic which is $z$ distance away from and parallel to the plane boundary. The relevant renormalized Hadamard function is given by Eq.~(\ref{eq:tpFunc}) with $\Delta z$ being replaced by $z+z'$. \vskip.25in \begin{figure} \begin{center} \leavevmode\epsfxsize=1.8in\epsfbox{fig5.eps} \end{center} \caption{ A light ray ( dashed line )propagates in the direction parallel to the plane boundary, starting $z$ distance away from it } \label{fig=fig5} \end{figure} Now suppose the geodesic starts at point $(t,a,0,z)$ and ends at point $(t',b,0,z)$, then the mean squared fluctuation in the geodesic interval function is \begin{equation} \langle \sigma_1^2 \rangle = {1\over 8}(b-a)^2 \int_{a}^{b} dz \int_{a}^{b} dz' \:\, G_{xxxx}^R(t,x,0,z,\,t',x',0,z). \, \ee Here $ G_{xxxx}^R(t,x,0,z,\,t',x',0,z) $ is also given by Eqs.~(\ref{eq:G1}-\ref{eq:G3}) but with a replacement of $nL$ by $2z$. Therefore \begin{eqnarray} &&\int_{a}^{b} dx \int_{a}^{b} dx'\, G_{xxxx}^{(1)R }(t,x,0,z,\,t',x',0,z) \nonumber\\ &&\quad ={2\over\pi^2 }\Bigg[ -\frac{\epsilon^4}{2(\epsilon^2+4)^2}-\frac{\epsilon^2}{(\epsilon^2+4)^2} \nonumber\\ &&\quad\quad \quad\quad+\frac{8\epsilon^5+ 32\epsilon^3+48\epsilon}{24(4+\epsilon^2)^{5/2}} \ln\left({\sqrt{4+\epsilon^2}+ \epsilon}\over{\sqrt{4+\epsilon^2}- \epsilon} \right)\Bigg]\,,\nonumber\\ \end{eqnarray} where $\epsilon =r/z $. Since the above expression is very complicated, we shall discuss two interesting special cases. One is when $r\gg z$, then we have \begin{equation} \langle \sigma^2 \rangle\approx {r^2\over 6\pi^2}\ln(r/z) \ee and \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r}=\sqrt{{\ln(r/ z)\over 6\pi^2} }\,. \ee This also blows up as $z$ approaches 0, however the functional dependence upon $z$ is different from that of Eq.~(\ref{eq:t1}). The other limit is when $r\ll z$. For this case, we find \begin{equation} \langle \sigma^2 \rangle\approx {r^8\over 720z^6\pi^2} \ee and \begin{equation} \Delta t={\sqrt{\langle \sigma_1^2 \rangle }\over r}={1\over 12\sqrt{5}\pi}\, \left({r\over z}\right)^3 \,. \ee \section{Summary and Discussion} In this paper, we have obtained general expressions, in the transverse tracefree gauge, for the vacuum graviton two-point function for various boundary conditions. These were used to study the lightcone fluctuations in flat spacetimes with a compactified spatial section and with a plane boundary. The mean squared fluctuations of the geodesic interval function and therefore the mean deviations from the classical propagation time have been obtained. In the case of a compactified spatial section, when the travel distance is less than the periodicity length, the fluctuation in the propagation time is less than the Planck time. In this limit, the effect is much larger for propagation in the periodicity direction than for propagation in the transverse direction. Thus the local lightcone fluctuations become anisotropic, reflecting the global structure of the spacetime. When the travel distance is large compared to the periodicity length, the fluctuation in travel time increases with the square root of the distance traveled for propagation in either direction, and the only difference lies in the proportionality constants. Here we have a possibility of having fluctuations larger than Planck scale by several orders of magnitude. In the case of a plane boundary, as light rays start closer and closer to the boundary, the lightcone fluctuations blow up as the square root of the logarithm of the starting distance both when light rays propagate perpendicularly and parallelly to the boundary. This is not as surprising as it might seem because the imposition of a fixed boundary can lead to singular expectation values of local observables, such as energy densities. However we expect this singularity to disappear if one treats the boundary as a quantum mechanical object with a nonzero position uncertainty \cite{Ford98}. It is also found that if the starting distance from the boundary is fixed, then the fluctuation in travel time grows as the square root of the logarithm of the distance traversed when this distance is large compared to the starting distance. In summary, we have demonstrated that in the linearized theory of quantum gravity, changes in the topology of flat spacetime produce lightcone fluctuations. These fluctuations are in general larger in the directions in which topology changes occur and are typically of the order of Planck scale, but they can get larger for path lengths large compared to the compactification scale. It is interesting to note that this effect could become significant in theories which postulate extra dimensions compactified on a very small scale. \begin{acknowledgments} We would like to thank Tom Roman for interesting discussions and X. Y. Zhong for help with graphics. This work was supported in part by the National Science Foundation under Grant PHY-9800965. \end{acknowledgments} \newpage \section*{Appendix} \setcounter{equation}{0} \renewcommand{\theequation}{A\arabic{equation}} \subsection{ Summation of graviton polarization tensors in the TT gauge} Let us introduce a triad of orthonormal vectors $({\bf e }_1({\bf k}) , {\bf e }_2({\bf k}) , {\bf e}_3({\bf k}) )$ with \begin{equation} {\bf e}_3({\bf k}) ={{\bf k}\over|{\bf k}|}=\hat{\bf k}, \ee the unit vector in the direction of propagation. The triad satisfies the orthonormality relation \begin{equation} {\bf e}_a({\bf k})\cdot{\bf e}_b({\bf k}) =\delta_{ab}, \quad \quad a,b =1,2,3. \ee This relation can be written, in terms of the components in the coordinate system characterizing the metric, as \begin{equation} e_a^i({\bf k}) e_b^i({\bf k}) =\delta_{ab}, \qquad a,b=1,2,3. \ee Here the Einstein summation convention is employed. We also have \begin{equation} e_a^i({\bf k}) e_a^j({\bf k}) =e^i_1e^j_1+e^i_2e^j_2+\hat k^i\hat k^j =\delta_{ij}, \qquad i,j= x,y.z. \ee Therefore, the two independent graviton polarization tensors in the TT gauge are given, in terms of the triad, by \begin{eqnarray} &&e^{ij}({\bf k},+)=e^i_1({\bf k}) \otimes e^j_1({\bf k}) -e^i_2({\bf k})\otimes e^j_2({\bf k})\,,\\ &&e^{ij}({\bf k},\times)=e^i_1({\bf k})\otimes e^j_2({\bf k}) +e^i_2({\bf k})\otimes e^j_1({\bf k})\,,\\ \end{eqnarray} where we have adopted the notation of Ref \cite{MTW}. Hence, \begin{eqnarray} \sum_{\lambda}\, e_{ij} ({{\bf k}, \lambda}) e_{kl} ({{\bf k}, \lambda})&&= e_{ij} ({{\bf k}, +}) e_{kl} ({{\bf k}, +})+ e_{ij} ({{\bf k}, \times}) e_{kl} ({{\bf k}, \times}) \qquad \qquad\nonumber\\ &&=e^i_1e^j_1e^k_1e^l_1-e^i_1e^j_1e^k_2e^l_2-e^i_2e^j_2e^k_1e^l_1+ e^i_2e^j_2e^k_2e^l_2\nonumber\\ &&\quad e^i_1e^j_2e^k_1e^l_2+e^i_1e^j_2e^k_2e^l_1+e^i_2e^j_1e^k_1e^l_2+ e^i_2e^j_1e^k_2e^l_1\nonumber\\ &&=(e^i_1e^k_1+e^i_2e^k_2)(e^j_1e^l_1+e^j_2e^l_2) +(e^i_1e^l_1+e^i_2e^l_2)(e^j_1e^k_1+e^j_2e^k_2)\nonumber\\ &&\quad -(e^i_1e^j_1+e^i_2e^j_2)(e^k_1e^l_1+e^k_2e^k_2)\nonumber\\ &&=(\delta^{ik}-\hat k^i\hat k^k)(\delta^{jl}-\hat k^j\hat k^l) +(\delta^{il}-\hat k^i\hat k^l)(\delta^{jk}-\hat k^j\hat k^k)\nonumber\\ &&\quad -(\delta^{ij}-\hat k^i\hat k^j)(\delta^{kl}-\hat k^k\hat k^l)\nonumber\\ &&=\delta_{ik}\delta_{jl} +\delta_{il}\delta_{jk}-\delta_{ij}\delta_{kl}+\hat k_i\hat k_j \hat k_k\hat k_l +\hat k_i \hat k_j \delta_{kl}+\nonumber\\ && \quad \hat k_k \hat k_l \delta_{ij}-\hat k_i \hat k_l \delta_{jk} -\hat k_i \hat k_k \delta_{jl}-\hat k_j \hat k_l \delta_{ik}-\hat k_j \hat k_k \delta_{il}\,. \label{eq:polsum} \end{eqnarray} This result can also be obtained as follows. Let us introduce a 4th-rank tensor \begin{equation} T^{ijkl}({\bf k})=\sum_{\lambda}\, e^{ij} ({{\bf k}, \lambda}) e^{kl} ({{\bf k}, \lambda})\,, \ee which has the following symmetry properties \begin{equation} T^{ijkl}=T^{jikl}=T^{ijlk}=T^{klij}. \ee However, the objects, which are at our disposal to construct $T^{ijkl}$, are only $k^i$ and $\delta^{ij}$, thus in general, we have \begin{eqnarray} T^{ijkl} &=& A\delta^{ij}\delta^{kl}+B\delta^{ik}\delta^{jl}+B\delta^{il}\delta^{jk}+ C(\hat k^i \hat k^j\delta^{kl}+\hat k^k\hat k^l\delta^{ij})\nonumber\\ && + D(\hat k^i \hat k^k\delta^{jl}+ \hat k^i \hat k^l\delta^{jk} + \hat k^j \hat k^l\delta^{ik}+ \hat k^j \hat k^k\delta^{il}) +E \hat k^i \hat k^j \hat k^k \hat k^l\,,\nonumber\\ \end{eqnarray} where $A,B,C,D,E$ are constants to be determined. This tensor is subject to the transversality condition \begin{equation} k_iT^{ijkl}=k_jT^{ijkl}=k_kT^{ijkl}=k_lT^{ijkl}=0\,, \ee and the trace-free condition \begin{equation} T^{iikl}=T^{ijkk}=0\,. \ee Applying these constraint conditions to $T^{ijkl}$ and solving the resulting equations leads to \begin{equation} a=d=-e=-c=-b\,. \ee Therefore $T^{ijkl}$ is the same as the right-hand side of Eq.~(\ref{eq:polsum}), apart for a multiplicative normalization constant which can be chosen to be unity. \subsection{ Vacuum graviton Hadamard function in the TT gauge} Here we evaluate the function $F_{ij}(x,x')$ and $H_{ijkl}(x,x')$ defined in Eqs.~(\ref{eq:Ffunc}) and (\ref{eq:Hfunc}), respectively. Once these functions are given, the graviton two point functions are easy to obtain. Define \begin{equation} R=\sqrt{(x-x')^2+(y-y')^2+(z-z')^2},\quad \Delta t=t-t',\quad k=|{\bf k}|=\omega \,. \ee Then, \begin{eqnarray} F_{ij}(x,x')&&={Re\over{(2\pi)^3}}\int\, d^3{\bf k}{k_ik_j\over{2 \omega^3}}e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega(t-t')} \nonumber\\ &&={Re\over{(2\pi)^3}}\partial_i\partial_j'\int_0^{\infty}\,{e^{-ik\Delta t}\over 2 k}\,dk \int_0^{\pi}\, d \theta\,\sin\theta e^{i k R \cos\theta}\int_0^{2\pi}\,d\phi\, \nonumber\\ &&={1\over{(2\pi)^2}}\partial_i\partial_j'\,{1\over R} \int_0^{\infty}\,{dk\over k^2} \sin kR\, \cos k\Delta t\,. \nonumber\\ \end{eqnarray} Because there is an infrared divergence in the above integral, we will introduce a regulator $\beta$ in the denominator of the integrand and then let $\beta$ approach 0 after the integration is performed. \begin{eqnarray} F_{ij}(x,x')&&={1\over{(2\pi)^2}}\partial_i\partial_j'\,\lim_{\beta\rightarrow 0}\,{1\over R } \int_0^{+\infty}\, {dk\over k^2+\beta^2} \sin kR\, \cos k\Delta t\, \nonumber\\ &&={1\over{(2\pi)^2}}\partial_i\partial_j'\,\lim_{\beta\rightarrow 0}\, f(\beta, R, \Delta t)\,. \nonumber\\ \end{eqnarray} Here we have used a integral in Ref.~ \cite{GR1} and defined \begin{eqnarray} f(\beta, R, \Delta t)&& = {1\over 4\beta R}\Bigg\{ e^{\beta(\Delta t-R)}{\rm Ei} [\beta(R-\Delta t)]+ e^{-\beta(\Delta t+R)}{\rm Ei} [\beta(R+\Delta t)]\nonumber\\ \quad &&- e^{\beta(\Delta t+R)}{\rm Ei} [-\beta(R+\Delta t)]-e^{\beta(R-\Delta t)}{\rm Ei} [\beta(\Delta t- R)]\Bigg\}\,.\nonumber\\ \end{eqnarray} Here ${\rm Ei(x)}$ is the exponential-integral function. Making use of the fact that, when $x$ is small, \begin{equation} {\rm Ei(x)}\approx \gamma +\ln|x|+x+{1\over 4}x^2+{1\over 18}x^3+O(x^4)\,, \ee where $\gamma$ is the Euler constant, and expanding $f$ around $\beta=0$ to the order of $\beta^2$, we get \begin{equation} f(\beta, R, \Delta t)\approx 1-\gamma -\ln \beta -{1\over 2}\ln ( R^2-\Delta t^2 ) -{\Delta t \over 4R}\ln \Bigg( \frac{R+\Delta t}{R-\Delta t}\Bigg)^2 +O(\beta^2)\,. \ee Taking the limit and keeping in mind that the constant terms ( with respect to $x$ and $ x'$ ) vanish under differentiation, we finally obtain \begin{equation} F_{ij}(x,x')=-{1\over{(2\pi)^2}}\partial_i\partial_j'\,\left[{1\over 2}\ln ( R^2-\Delta t^2 ) +{\Delta t \over 4R}\ln \Bigg( \frac{R+\Delta t}{R-\Delta t}\Bigg)^2 \right]\,. \ee Now let us turn our attention to $H_{ijkl}(x,x')$. We have, proceeding with similar steps as we did for $F_{ij}(x,x')$, \begin{eqnarray} H_{ijkl}(x,x')&&={Re\over{(2\pi)^3}}\int\, d^3{\bf k}{k_i k_j k_k k_l\over{2 \omega^5}}e^{i{\bf k} \cdot({\bf x}-{\bf x'})}e^{-i\omega\Delta t}\nonumber\\ &&=-{1\over{(2\pi)^2}}\partial_i\partial_j'\partial_k\partial_l'\, \lim_{\beta\rightarrow 0}\,{1\over 2\beta R }{\partial\over \partial\beta} \int_0^{+\infty}\, {dk\over k^2+\beta^2} \sin kR\, \cos k\Delta t\, \nonumber\\ &&=-{1\over{(2\pi)^2}}\partial_i\partial_j'\partial_k\partial_l'\,\lim_{\beta\rightarrow 0}\, {1\over 2 \beta } {\partial\over \partial\beta} f(\beta, R, \Delta t)\,. \nonumber\\ \end{eqnarray} Now expand ${1\over 2 \beta } {\partial\over \partial\beta} f(\beta, R, \Delta t) $ to order $\beta^2$ to find \begin{eqnarray} {1\over 2 \beta } {\partial\over \partial\beta} f(\beta, R, \Delta t)\,&&=-{1\over 2\beta}-{1\over 3} \left[ (\ln \beta +\gamma-1)R^2+3(\ln \beta +\gamma-1)\Delta t^2\right]\nonumber\\ \quad &&-{1\over 12 R}\left[ (R+\Delta t)^3\ln|R+\Delta t| + (R-\Delta t)^3\ln|R-\Delta t|\right]\,.\nonumber\\ \end{eqnarray} Plugging this result into Eq.~(A22) and noting that only terms higher than quadratic in $R$ contribute after the differentiation, we obtain \begin{eqnarray} H_{ijkl}(x,x')&=&{1\over 48\pi^2}\partial_i\partial_j'\partial_k\partial_l'\, \Bigg[ (R^2+3\Delta t^2)\ln (R^2-\Delta t^2)^2\nonumber\\ \quad && + \left(3R\Delta t+{\Delta t^3\over R}\right)\ln \Bigg( \frac{R+\Delta t}{R-\Delta t}\Bigg)^2\Bigg]\,. \nonumber\\ \end{eqnarray} For convenience, we give the explicit forms for $G_{xxxx}^{(1)}$ and $G_{zzzz}^{(1)}$ here: \begin{eqnarray} G_{xxxx}^{(1)}(x,x')\,&&=2\left(D^{(1)}(x,x')-2F_{xx}(x,x')+H_{xxxx}(x,x')\right)\nonumber\\ &&={1\over 12\pi^2R^8\sigma^2}\Bigg\{(\Delta x^2-\Delta t^2)(16\Delta x^6-24\Delta x^4\Delta t^2) -3\Delta t^2 R_t^6\nonumber\\ &&\quad +(9\Delta t^4+69\Delta x^2\Delta t^2+16\Delta x ^4)R_t^4+(-72\Delta x^2\Delta t^4+32\Delta x^4\Delta t^2+32\Delta x^6) R_t^2\Bigg\} \nonumber\\ &&\quad-{\Delta t\over{16\pi^2R^9}}\ln\left(\frac{R+\Delta t}{R-\Delta t}\right)^2 \Biggl[-R_t^6-(3\Delta t^2 +9\Delta x^2)R_t^4\nonumber\\ &&\quad +24\Delta t^2\Delta x^2R_t^2-8\Delta x^4\Delta t^2 +8\Delta x^6\Biggr]\,,\nonumber\\ \label{eq:tpFunc} \end{eqnarray} where \begin{eqnarray} &&R_t^2=\Delta y^2+\Delta z^2\\ &&\Delta x =x-x'\\ &&R^2=R_t^2 +\Delta x ^2=\Delta x ^2 +\Delta y^2+\Delta z^2\\ &&\sigma^2=R^2-\Delta t ^2=\Delta x ^2 +\Delta y^2+\Delta z^2-\Delta t ^2\\ \end{eqnarray} To get $G_{zzzz}^{(1)}(x,x')$, all we need to do is to replace $R_t^2 $ in Eq~(\ref{eq:tpFunc}) by $R_t^2=\Delta y^2+\Delta x^2$.
\section{Introduction} In the last 20 years the most widely used technique to analyze X--ray images has been the so-called ``sliding box'' technique: a box of fixed size is passed across an image and sources are identified as enhancements of the total photon counts in the box against the background, which is independently estimated via local or global measurements. This technique has been used in the past to build source catalogs from imaging X-ray missions, such as Einstein (Harris 1990), EXOSAT (Giommi et al. 1991) and ROSAT/PSPC (White, Giommi \& Angelini 1994). A different approach has been used in the compilation of the ROSATSRC \markcite{zim94} (Zimmermann 1994) and the Rosat All Sky Survey BSC \markcite{vog96} (Voges et al. 1996) catalogs, where maximum likelihood filters were utilized for source detection and characterization. Although these catalogs have proven to be extremely useful, they also have shown important limitations. In deep confusion limited images, especially crowded fields, or in the presence of very bright or extended sources (such as clusters of galaxies and supernova remnants), these techniques often tend to blend multiple sources into a single one, to split a bright source in multiple detections and/or to misidentify small background fluctuations as real sources. In addition, these catalogs provide a poor estimate (if any) of the intrinsic source size and the distinction between extended and point-like emission is hardly possible. This means that a visual check ``source by source'' is required in most cases in order to ``clean'' the source catalog. An improvement of these techniques has been presented by Vikhlinin et al. 1995, who substituted the sliding box with a matched filter that reproduces at best the profile of point-like sources. This method is optimal for detection of point-like sources, but tends to miss faint sources that are spatially resolved by the detector. The next generation of X--ray missions, with very high sensitivity and spatial resolution, will produce very deep images with a very high surface density of X--ray sources (up to $1000/{\rm deg{^2}}$) and a level of morphological detail comparable to that of optical images. Therefore, it is important to develop new and more refined techniques for source detection and characterization to fully exploit the scientific content of these vast datasets. These problems and considerations led to the development of a new generation of detection algorithms in which the square window of the sliding box technique is replaced with a family of filters of different sizes with particular morphological and functional properties: the Wavelet Transform (hereafter WT). The WT is a mathematical tool capable of decomposing an image in a set of sub-images, each representing the image features at a given scale. Sources with sizes ranging from the intrinsic instrumental resolution up to a significant fraction of the whole image can be efficiently extracted at the WT scale that better matches the source size. Moreover WT-based detection algorithms are insensitive to large scale background fluctuations provided that the source size is different from this scale, a typical situation in high energy images. The application of WT-based algorithms to the automatic detection and characterization of sources in X--ray images was first used by Rosati et al.\markcite{ros93,ros95a,ros95b} 1993, 1995 (see also Rosati 1995), who exploited the multi-scale capabilities of this method to compile a catalog of extended sources in ROSAT-PSPC deep pointed observations. This work led to the compilation of the ROSAT Deep Cluster Survey \markcite{ros95b,ros98} (RDCS; Rosati et al. 1995, 1998), a deep survey of galaxy clusters over $\sim 50$ square degrees. More recently, WT based techniques have been developed and applied by other groups for the analysis of ROSAT, mainly PSPC, images \markcite{vik95,gre95,har96,pis97,dam97a,dam97b,pis98} (Grebenev et al. 1995; Harnden et al. 1996; Pislar et al. 1997; Damiani et al. 1997a,b; Vikhlinin et al. 1998), even if a complete catalog of WT detected sources has not been published to date. In this and in the accompanying paper \markcite{cam98} (Campana et al. 1999; Paper II hereafter) we present a multi-scale wavelet-based algorithm which generalises and combines those used by Rosati et al. 1995 and Lazzati et al. 1998, and its application to the complete archival data of the ROSAT High Resolution Imager (HRI). The paper is organized as follows: in Sect.~\ref{algosec}~and~\ref{charasec} we describe how the detection and characterization algorithm works, while in Sect.~\ref{simulsec} we test its performances with simulations. Finally, in Sect.~\ref{disc} we summarize our results and briefly discuss the applications of our algorithm. \section{The detection algorithm} \label{algosec} The basic steps of the detection algorithm we developed are summarized in Figure~\ref{blops}, while a more detailed description of single steps is given in the following sections. First, a wavelet convolution is performed on the X--ray image and an average background value is estimated with a $\sigma$-clipping algorithm (see Sect.~\ref{backsec}). Given the measured background and referring to the simulations described in Sect.~\ref{simulsec}, a detection threshold is computed constraining the total average number of spurious detections in the whole image. The source candidates are selected as significant (above the threshold) peaks in each independent wavelet sub-image and a complete catalog is obtained cross-correlating the single-scale catalogs to eliminate multiple-scale detections of the same source and the large-scale blending of nearby small-scale sources. At this stage, the catalog merely consists of a coarse position and a rough estimate of the count rate and size of the sources. The second part of our algorithm (see lower box of Figure~\ref{blops}) deals with the refinement of these quantities along with their uncertainties and leads to the complete source characterization. To this aim, a model source with a Gaussian point spread function (hereafter PSF) is fitted in wavelet space with a multi-dimensional procedure which takes into account the behavior of the wavelet coefficients both in space and scale. Up to 20 neighbor sources can be fitted simultaneously in the deblending process, thus allowing the characterization of faint objects located near bright sources \markcite{laz98} (see also Lazzati et al. 1998). As a final result, the algorithm produces - in a fully automated way - a catalog of sources with position, count rate and apparent size along with corresponding uncertainties. For each source the associated probability of spurious detection is also derived. \subsection{The wavelet transform} The wavelet transform of a real function $f$ is the convolution product between $f$ and a class of analyzing wavelets $\psi_a(x,y)$, all derived by dilation from a single function $\psi_1(x,y)$, called mother wavelet: \begin{equation} \psi_a(x,\,y) = \frac 1a \psi_1(\frac xa,\, \frac ya). \end{equation} \noindent Any function $g(x,y)$ can be used as the mother for a bidimensional wavelet transform if it satisfies the following conditions: \begin{eqnarray} \int_{\mathbb{R}^2} |g(x,y)|^2 \, dx \, dy &<& \infty \nonumber \\ \int_{\mathbb{R}^2} g(x,y) \, dx \, dy &<& \infty \end{eqnarray} The above properties assure the existence of the transform of any square integrable function (as astronomical images are) and, more importantly, the fact that, whatever the function $f$ and the scale $a$ of the transform, the mean value of the transformed image is null. This provide an objective and automatic background subtraction in detection applications of the WT. For a detailed and rigorous analytical description of the wavelet operators, we refer the reader to the extensive reviews published in the astronomical literature \markcite{gre95,sle94} (e.g. Grebenev et al. 1995; Slezak et al. 1994) or to more general discussions \markcite{mal92,you93} (Mallat \& Hwang 1992; Young 1993) and references therein. In this work we adopt a discrete approximation of the wavelet operator \markcite{gre95,laz98} (see, e.g. Grebenev et al. 1995; Vikhlinin et al. 1998; Lazzati et al. 1998), which keeps all the properties described above provided that the integrals are replaced with discrete summations. This allows faster computations and a reduction of the cross talks between adjacent pixels and scales. The continuous approach, which has been adopted by - e.g. - \markcite{dam97a,dam97b} Damiani et al. 1997a,b, allows the calculation of the transform at whatever scale and position. In this scheme, source characterization proceeds through an iterative refinement of the position and scale at which the transform is calculated. Our WT computation is based on the multi-resolution theory \markcite{far92} (Farge 1992) and the ``\`a trous'' algorithm \markcite{bij92} (Bijaoui \& Giudicelli 1992). We use the `mother' developed by \markcite{sle95} Slezak et al. 1995, which can be analytically approximated with the difference of two Gaussians: \begin{eqnarray} \psi_1(r) &=& A \, \exp \left[-{\frac{r^2}{2\sigma_A^2}}\right] - B \, \exp\left[-{\frac{r^2}{2\sigma_B^2}}\right] \label{pareq} \\ A &\simeq& 0.597; \;\;\;\;\; \sigma_A \simeq 0.687 \nonumber \\ B &\simeq& 0.264; \;\;\;\;\; \sigma_B \simeq 1.030 \nonumber \end{eqnarray} \noindent where the $(x,y)$ dependence has been turned into radial due to the symmetry of the mother. This symmetry gives us a new property of the WT operator: it is insensitive to first order derivatives of the transformed function, i.e. the transform of a plane is constantly equal to zero. This means that smoothly varying backgrounds will not affect the significance of the detections. The multi-resolution technique provides us with a sampling on scales logarithmically spaced by a factor of two. According to the wavelet theory (see e.g. \markcite{you93} Young 1993) this provides the best compromise between completeness and correlation (i.e. redundancies in the wavelet coefficients): a finer sampling would cause an oversampling while a coarser sampling would produce a loss of information. We hence have: $$a_j = 2^j; \qquad j\in \mathbb{N}$$ \noindent The sampling in the $x$ and $y$ is performed pixel by pixel, at any scale. This produces an heavy oversampling of the transform, especially at the largest scales, but it does not affect the detection. In the second part of the algorithm, when sources are characterized, a decimation process is applied in order to minimize this oversampling (see Sect.~\ref{decimasec}). \subsection{Detection thresholds} \label{threshsec} Due to the structure of the wavelet operator, the statistics of coefficients in the WT space cannot in general be handled analytically. If, however, the function $f$ is a white Gaussian noise with mean $\bar{d}$ and dispersion $\sigma_d$, we have that the individual WT coefficients are Gaussian distributed according to the law: \begin{eqnarray} p(WT) &=& \frac {1}{\sqrt{2\pi\sigma^2}} \, \hbox{exp} \Big (-\frac{WT^2} {2\sigma^2} \Big ) \label{wtprob} \\ \sigma &=& \sqrt{\sigma_d^2 \; \int |\psi|^2} \label{flatsig} \end{eqnarray} \noindent Unfortunately, this is not the case for high energy astronomical images that can be approximated by a white Poisson noise of constant mean (the background\footnote{Here we neglect the vignetting and the effects described in Paper II.}), on which sources are overlaid. Since our goal is to assess the probability that the WT peaks are random background realizations, we can neglect the statistics of the sources and concentrate on the transform of pure background. It can be shown (see, e.g. \markcite{gre95} Grebenev et al. 1995) that Equation~\ref{flatsig} still holds for a Poisson noise, with $$\sigma = <WT^2> - <WT>^2 = <WT^2>; \quad \sigma_d = \sqrt{ \bar d}$$ \noindent however, the probability distribution given by Equation~\ref{wtprob} is not followed anymore. Moreover, even if we knew exactly the distribution of the single coefficients, the WT non-orthogonality would prevent us from computing the completeness and contamination of the catalog. Thus, to overcome these problems we must rely on Monte Carlo simulations. For a set of mean background values between $10^{-4}$ and $10$ counts per pixel we simulated 5000 realizations of random fields, $512\times512$ pixels each. Then we computed the WT and its standard deviation with Equation~\ref{flatsig} and counted all local maxima that exceeded a given signal-to-noise threshold (i.e. the ratio of the wavelet transform and the error given in Equation~\ref{flatsig}). Since no source has been simulated, any peak in the transformed image is spurious. Repeating this procedure with different thresholds for each scale $a$ of the transform, we have built the integrated probability function of finding a random WT peak exceeding a given signal-to-noise. Figure~\ref{soglieps} shows the typical output of a set of simulations with a background of 0.3 counts per pixel. The number of WT peaks that exceeds a given signal-to-noise in the whole field is plotted versus the signal to noise itself for the different scales. To limit the number of spurious detections to, e.g., one in ten fields, we draw a horizontal line corresponding to 0.1 sources and, scale by scale, we set the threshold at the signal to noise where the horizontal line crosses the corresponding curve. Combining the results of all simulations we obtain the integrated probability function $P$ which, given the number of pixels of the search, the desired signal to noise threshold and the mean value of the background counts, returns the mean expected number of spurious detections. In the algorithm, this function must be inverted. In fact, when we run the WT algorithm on an image, we fix the expected contamination (the number of spurious detections in the whole image) and we compute - as a function of the scale $a$ - the threshold to be used. Note that, since the number of independent points is proportional to the square of the inverse of the scale, the threshold applied on smaller scales will be higher than that on larger scales (see Figure~\ref{soglieps}). In principle, the whole set of scales provided by the wavelet transform can be used in the detection procedure. However, in the analysis of high energy images - as those of the ROSAT HRI (see Paper II) - we have a lower scale for the sources set by the instrumental resolution. The analysis on very large scales, say half the size of the image, is strongly contaminated by edge effects and sensitive to residuals from the image reduction. As a general rule, the best scales on which the analysis can be performed are the ones which closely match the instrumental PSF. For these reasons, in the following of this paper and in the analysis of HRI images (Paper II), we will use only the following scales: \begin{equation} a = 5; \, 10; \, 20; \, 40 \quad \hbox{pixels} \label{scaleq} \end{equation} \noindent where the scale $a$ roughly measures the FWHM of the positive central part of the wavelet filter in pixels. \subsection{Different scales cross-identification} \label{crosec} Since the source detection is performed on several scales and given the non-orthogonality of the wavelet operator, a cross-correlation of the single-scale catalogs must be performed. This allows us to get rid of multiple identifications of the same source at different scales and, more importantly, to discriminate true extended sources from a blending of nearby point-like sources. This latter distinction is not trivial, since true extended sources often produce apparently significant peaks in the lower scales, while nearby sources are inevitably blended at larger scales. First, we eliminate from the catalog multiple scale detection of bright sources on the basis of position and signal-to-noise: among all sources with compatible positions only the most significant (i.e. with the highest signal-to-noise ratio) is held. Two sources detected at different wavelet scales but with nearby positions are matched in a single source if their distance is lower than the scale at which the most significant has been detected. A more difficult problem is dealing with the merging of nearby sources at large scales. Two or more nearby point-like sources are usually detected as a single extended source at large scales. However, it is also possible that a real extended feature is overlaid on them (a typical situation in supernova remnants and in nearby clusters of galaxies). Our cross-correlation scheme held the large scale detection as a true source only if its significance (signal-to-noise) is larger than that of all lower-scale sources overlaid on it. This scheme slightly favors point-like sources. The worse situation is provided by a group of weak and close sources, which are detected only at large scales as a single extended feature. This is an intrinsic limitation of the detector resolution that cannot be solved by a detection algorithm, however refined (see, e.g. Damiani et al. 1997a). Rosati et al. 1995 describe a different scheme, in which all sources are detected and characterized for each scale separately. A subsequent step consists in the cross-correlation of different scale catalogs to get rid of multiple and false detections by means of positional coincidence and fitted count-rates. Damiani et al. 1997a cross-identify only sources detected at consecutive scales and match sources if their distance is less than the scale size corresponding to the maximum signal-to-noise (or 1.5 times the detector PSF). Albeit simple, our scheme has been shown to be powerful and robust in the analysis of ROSAT HRI fields, that can be very crowded of point sources (see Paper II, in particular Figure 10). \section{Source characterization} \label{charasec} After the whole procedure described in last section, we are left with a list of sources with a rough determination of position (the center of the pixel with higher WT coefficient), source size (the scale of the WT where the signal to noise is maximized) and total number of photons (the value of the maximum WT coefficient). Several schemes have been proposed in the literature to refine these estimates in wavelet space; these can be divided in two main groups: fitting the multi-scale profile \markcite{gre95,ros95b} (e.g. Grebenev et al. 1995; Rosati et al. 1995) or searching for the best position and scale of the source recomputing the WT in a narrower grid both in space and in scale \markcite{dam97a,dam97b} (Damiani et al. 1997a,b). Our approach is to fit source profiles both in space \markcite{ros95b} (as in Rosati et al. 1995) and in scale (as in \markcite{gre95} Grebenev et al. 1995). This allows us to refine simultaneously the position, size and flux of the sources without the need of weight summing parameters derived at different scales. The narrower grid method should in principle be more robust, since it does not involve multi-parameter fitting procedures that could fail when run automatically. However, such a method needs analytic WT computations (instead of our numeric method) and makes it difficult to deal with the strong cross terms in WT space in presence of nearby sources and, more generally, in crowded fields. \subsection{Decimation technique} \label{decimasec} To apply standard routines ($\chi^2$ minimization) to the source characterization in wavelet space, we must first assure that the hypothesis of independence between input data is appropriate. As stated above, this is not the case, mostly at larger scales. Although \markcite{ros95a} Rosati 1995 has shown that neglecting the correlations between adjacent data does not significantly compromise the result of the fit, the uncertainties on extracted parameters are highly underestimated and the oversampling in $x,y,a$ lengthens considerably the computing time. To obtain a fast, robust and unbiased estimate of the source parameters and related errors, we have developed a decimation technique that, taking into account the correlation properties of the WT, extracts a subset of quasi-independent coefficients from the whole transform (see also \markcite{laz98} Lazzati et al. 1998). Before applying the decimation, we first select a subregion of the transform around the source position. For each detected source, named $\tilde a$ the most relevant scale of the detection, a region of radius $4 \, \tilde a$ centered on the source position is extracted from the whole transform and the scale range is reduced to the set $\{\tilde a/2,\, \tilde a,\, 2\, \tilde a\}$. Scales smaller than $\tilde a/2$ are dominated by noise, while those larger than $2 \, \tilde a$ suffer from the contamination from adjacent sources. To extract a subset of wavelet coefficients which is almost independent, the minimum distance between two coefficients at a given scale must be almost equal to the correlation length of the transform at this scale, which is proportional to the scale itself. From the analysis of the autocorrelation function of the transform of a pure white noise, we find that the correlation length is given by: $c_l(a) \sim 1.7 \, a$ (see also \markcite{laz96} Lazzati 1996). At the scales mentioned above we should have, in principle, 69, 17 and 4 orthogonal wavelet coefficients, respectively. Since the most significant coefficient at all scales is the coefficient centered on the source position, we keep it in the decimated set. We finally obtain 61, 17 and 5 coefficients for the three scales. The position of this coefficients for the case $\tilde a = 5$ is plotted in Figure~\ref{decima}. These extracted coefficients provide us with a small set of almost orthogonal\footnote{It is not possible to extract a set of rigorously orthogonal coefficients, but this decimation proves to be successful {\it a posteriori} since standard fitting routines produce reliable results.} and Gaussian distributed points. \subsection{Source fitting} The model we fit to the decimated set of WT coefficients is the transform of a bidimensional Gaussian. This transform can be computed analytically and can be written as: \begin{eqnarray} \tilde f (x,y) &=& 2\pi I_0 \bigg [ A \sigma_A^2 \, G\Big(x,y,x',y', \sqrt{\sigma^2 + 2^{2j}\sigma_A^2} \Big) - \nonumber \\ &-& B \sigma_B^2 \, G\Big(x,y,x',y', \sqrt{\sigma^2 + 2^{2j}\sigma_B^2} \Big) \bigg] \label{treq} \end{eqnarray} \noindent where $I_0$ is the integral of the source, $\sigma$ its width, $A$, $B$, $\sigma_A$ and $\sigma_B$ are given in Equation~\ref{pareq} and $G$ is defined as: \begin{equation} G(x,y,x',y',\sigma) = \frac1{2\pi\sigma^2}\exp \bigg[ -\frac{(x-x')^2+(y-y')^2}{2\sigma^2} \bigg] \end{equation} \noindent The output of the procedure is hence a set of values $\{x', y', I_0, \sigma \}$ along with their errors. Since our fitting procedure involves scales larger than the natural source scale, it is possible that the WT coefficients are affected by the presence of nearby sources. To deal with this problem, multiple source fitting (up to 20 simultaneous sources) can be performed. Thanks to the linearity of the WT operator, the transform of $N$ sources in a small area is: $$\tilde f (x,y) = \sum_{i=1}^N \tilde f_i (x,y)$$ \noindent where $\tilde f_i (x,y)$ is given by Equation~\ref{treq}. In practice, since it would be meaningless to fit $20 \times 4 = 80$ parameters with only 83 (i.e. $5+17+61$) data and since weak sources could hardly affect the fit of a strong source, the procedure is structured as follows. For each source, the algorithm searches for all sources within ten times its estimated width. In this set of sources, the parameters of those already fitted are freezed while the total counts and sizes of all the other sources are free in the fitting routine. When all sources have been fitted once, the procedure is repeated (improved fit in Figure~\ref{blops}), but this time all nearby sources are held fixed. This multi-source fitting allows a more precise determination of the source fluxes. This is particularly true in the case of very nearby sources where the influence of the negative wings of one source modifies the value of the other source transform (see e.g. \markcite{laz98} Lazzati et al. 1998). \subsubsection{Error determination} If the number of background and source counts is sufficient ($\gtrsim 5\times 10^{-2}$~counts per pixel), a reliable estimate of the parameter uncertainties can be directly obtained from the covariance matrix (see Sect.~\ref{simulsec} for more details). However, a more detailed computation of the error on WT coefficients is needed. In fact, Equations~\ref{wtprob} and~\ref{flatsig} are no longer valid to estimate the WT fluctuations in presence of sources (i.e. non-flat components). A more general form for Equation~\ref{flatsig} is given in \markcite{gre95} Grebenev et al. 1995: \begin{equation} \sigma(x,y) = \sqrt{ f(x,y) \ast |\psi|^2} \label{nfsig} \end{equation} \noindent where $\ast$ represents the convolution product in $\mathbb{R}^2$ and $f(x,y)$ is the function to be transformed. Poisson statistics for $f$ has been assumed. Note that Equation~\ref{nfsig} is consistent with the results given above when $f=\hbox{const.}$ (cf. Equations~\ref{wtprob},~\ref{flatsig}). When dealing with very low backgrounds ($\lesssim 5\times 10^{-2}$~counts per pixel), in particular at the lowest scales, the error determination suffers from the strong departure of the WT statistics from the Gaussian case. As a result, errors are generally underestimated. In this case a precise determination in WT space would imply a fitting procedure based on WT statistics which, as stated above, cannot be handled analytically. However, from basic statistics, we can obtain reliable estimates of errors with an {\it a priori} method. Let's consider position $x'$ and $y'$. If we have a Gaussian shaped spatial distribution of $N$ photons with a width $\Delta$, the precision on the determination of the centroid is limited by: \begin{equation} \sigma_{x'} \simeq \sigma_{y'} \simeq \frac{\Delta}{\sqrt{N-1}} \label{xunc} \end{equation} \noindent while the error on total number of counts will be given by Poisson statistics: \begin{equation} \sigma_{N} \simeq \sqrt{N} \label{func} \end{equation} \noindent The intrinsic limit on the width estimate can be expressed through the rms scatter of photon positions: \begin{equation} \sigma_\Delta \simeq 0.5 \, \frac{\Delta}{\sqrt{N-1}} \label{sunc} \end{equation} The final error on fitted parameters is set to be the maximum value between that estimated from the covariance matrix and the above equations. \section{Testing the algorithm} \label{simulsec} To test the reliability of the source characterization and of the parameter uncertainties, we have simulated a set of $21$ fields with $64$ sources each, for a set of backgrounds ranging from $10^{-3}$ to $10^{-1}$~counts per pixel. Hence, for each background value we have 1344 synthetic sources. The source positions are fixed to an equally spaced $8\times8$ grid in order to test the accuracy of the fit without edge problems and source confusion (see also Paper II for a complete discussion of this problem). To reproduce at best an astronomical image, the total counts of sources have been simulated according to an Euclidean ${\rm Log}\, N - {\rm Log}\, S$, while all sources have been assumed point-like with a FWHM~$\simeq 6$~pixels. We have then applied the full procedure described above, including background determination and source detection with a threshold of 0.1 spurious sources per simulated field. \subsection{Background determination} \label{backsec} The estimate of the mean background value is carried out with a $\sigma$-clipping method. This procedure relies on the hypothesis that the statistics of the image is dominated by two independent distributions: the background counts and the source counts. If the average values of these distributions are strongly different (which is very often the case when dealing with high energy astronomical images) the contribution of sources can be iteratively erased and a robust mean background value measured. To reach this goal the mean and standard deviation of the whole image are calculated and pixels with a number of counts that exceeds the mean value by more than 3 standard deviations are flagged. The process is repeated iteratively on unflagged pixels until the mean and the standard deviation values converge. Since the $3 \, \sigma$-clipping implies Gaussian statistics, the images are binned to obtain an initial mean number of counts per pixel of at least 10. \subsection{Results} \label{risultati} These simulations have revealed that as long as the distribution of WT coefficients is well approximated by a Gaussian function, the procedure works very well. With small background values - below $\sim 5\times 10^{-2}$ counts per pixel - the distribution in the lowest scales of WT space becomes increasingly Poissonian and the covariance matrix gives underestimated errors, even if the parameter evaluation remains reliable. In this case, the uncertainties given by Equations~\ref{xunc},~\ref{func}~and~\ref{sunc} turn out to be more accurate than those obtained in the fitting procedure. Sources with even five input photons (the minimum input value) are detected in the lowest background simulations. This is not a surprising result because, as pointed out by Damiani et al. 1997a, in principle it is possible to detect sources with 3 photons only. Figures~\ref{s1ps},~\ref{s2ps}~and~\ref{s3ps} show the result of these simulations. In all figures, the first panel shows the absolute discrepancy between the input and output parameters (position, total counts and width, respectively) versus the input counts. In the right panels the distribution of the absolute discrepancy between the input and output parameters, divided by the their estimated errors, are compared with a normal Gaussian distribution. Asterisks and dashed lines refer to the lower background ($10^{-3}$~counts per pixel) simulations, while circles and solid lines to a higher background case ($10^{-1}$~counts per pixel). For low background simulations (dashed lines in Figures~\ref{s1ps},~\ref{s2ps}~and~\ref{s3ps}) the correlation matrix gives underestimated errors and Equation~\ref{xunc}, \ref{func} and \ref{sunc} are always used. A Kolmogorov-Smirnov (KS) test on the whole set of simulated sources confirms the visual impression that dashed histograms are not incompatible with the Gaussian solid line in the figures. For high background simulations (solid histograms) the correlation matrix is commonly used. Considering the whole set of simulated sources, errors are slightly underestimated (20\%, especially in Figure~\ref{s1ps}). However, a closer inspection reveals that when only bright ($>50$~cts) sources are selected, the estimated errors are Gaussian distributed. We note that the overestimation of fluxes near the detection threshold is common, due to the fact that sources on top of positive background fluctuations are preferentially selected. Figure~\ref{s1ps} shows that source positions are determined with an accuracy which is of the order of the source half width in the low signal-to-noise regime. A similar result is obtained for the total counts (Figure~\ref{s2ps}). Only for one source out of $2500$ the total counts evaluation has failed by more than a factor of two. This faint source, whose flux and size measurements are affected by a large error, lays in the close vicinity of one of the four significant positive background fluctuations expected (40 simulated fields with 0.1 spurious sources each). The algorithm merged the true source and the fluctuation in a higher flux extended feature. For what concerns the measurement of the source size, Figure~\ref{s3ps} shows that particularly in the lower background simulations the width of the source is underestimated. This is due to the fact that, since the vast majority of sources in our simulations are at the limit of detection, those that randomly happen to be more compact have a higher signal-to-noise ratio and are hence preferentially selected by the algorithm against those with lower surface brightness. This bias is present only for weak sources in very low background images and would cause the misidentification of extended sources as point-like and not vice-versa. The shaded region in the right panel of Figure~\ref{s3ps} shows how this bias is considerably reduced when only sources with more than 100 counts are considered. \section{Conclusions} \label{disc} We have presented a new detection algorithm based on the wavelet transform for a multi-scale analysis of astronomical X--ray images which is suited for the detection and the characterization of both point-like and extended sources. Given an image in the X-ray band, the algorithm produces a catalog of sources characterized by position, count rate and size along with associated errors. The main difference of our technique with respect to other WT-based algorithms lies in the source characterization. Our algorithm uses, for the first time, a fitting procedure that models the source in WT-space by taking into account the spatial and multi-scale behavior simultaneously. Moreover, this procedure is performed on a decimated set of wavelet coefficients to drastically reduce the cross-talk between adjacent pixels and scales. Future high energy missions equipped with CCD detectors, such as XMM, AXAF and JET-X, will combine a high effective area with a small instrumental background. In these cases, current wavelet implementations which rely on Gaussian statistics for the source characterization may not lead to optimal results. The problem of crowded fields has been dealt with a multi-source characterization procedure, where up to 20 sources are fitted simultaneously. In very low background images, this fitting procedure produces reliable parameter estimates but underestimated errors. In these cases we have developed a different error evaluation procedure which relies on basic statistics and which proves to be robust even with vanishing backgrounds. The use of wavelet-based detection algorithms for the data analysis of the new generation of X--ray missions will provide an accurate, fast and user friendly source detection software. A first application of this algorithm to high resolution X-ray images obtained with the ROSAT HRI is presented in the accompanying paper \markcite{cam98} (Campana et al. 1999) and the analysis of the full set of archival HRI pointings is underway for the construction of a complete wavelet-based catalog of ROSAT HRI X-ray sources. \acknowledgements{This work has been supported through ASI and CNAA grants.}
\section{Introduction.} Two-dimensional turbulence has very peculiar properties compared to 3d turbulence, see eg. ref.\cite{frisch,review} for references. As first pointed out by Kraichnan in a remarkable paper \cite{kraich}, these open the possibility for quite different scenario for the behaviors of turbulent flows in 2d and 3d: if energy and enstrophy density are injected at a scale $L_i$, with respective rate $\overline \epsilon$ and $\overline \epsilon_w\simeq \overline \epsilon L_i^{-2}$, the 2d turbulent systems should react such that the energy flows toward the large scales and the enstrophy towards the small scales. One usually refers to the infrared energy flow as the inverse cascade and to the ultraviolet enstrophy flow as the direct cascade. In the (IR) inverse cascade, scaling arguments lead to Kolmogorov's spectrum, with $E(k)\sim \overline \epsilon^{2/3}\,k^{-5/3}$ for the energy and $(\delta} \def\De{\Delta u) \sim (\overline \epsilon r)^{1/3}$ for the variation of the velocity on scale $r$. In the (UV) direct cascade, scaling arguments give Kraichnan's spectrum with $E(k)\sim \overline \epsilon_w^{2/3}\, k^{-3}$ for the energy and $(\delta} \def\De{\Delta u) \sim (\overline \epsilon_w r^3)^{1/3}$ for the velocity variation. The aim of this Letter is to discuss a few possible scenario for the influence of friction on the direct enstrophy cascade of 2d turbulence. The picture which emerges is that at small scales the velocity field has a smooth gaussian component, with amplitude diverging in the frictionless limit, supplemented by possible anomalous corrections whose characters depend whether there is entrophy dissipation or not in the inviscid limit. We also point out that friction provides a way to regularize amplitudes in the inviscid limit. \medskip {\bf A few experimental facts.} Two-dimensional turbulence has recently been observed in remarkable experiments providing new informations on these peculiar characteristics. See refs.\cite{tabel,tabis} for a precise description of the experimental set-up. The turbulent flow takes place in a square cell such that the injection length is $L_i\simeq 10\ cm$. The bottom of the cell induces friction with a coefficient $\tau \simeq 25\ s$. The dissipation length $l_d$ is of order $1\ mm$ and the UV friction length $l_f$ of order $0.5\ cm$. ($\tau,\ l_d,\ l_f$ are defined below). The enstrophy transfer rate determined from the three point structure function has been evaluated in \cite{tabis} as $\eta_w\simeq 0.4\ s^{-3}$. The direct cascade is observed at intermediate scales between $1\ cm$ and $10\ cm$. Besides the absence of any experimentally significant deviations from Kraichnan's scaling, two remarkable facts have been observed: (i) the vorticity structure functions are almost constant on the inertial range with the two-point function of order $10\ s^{-2}$; and (ii) the vorticity probability distribution function is almost symmetric and almost, but not quite, gaussian. \medskip {\bf A model system.} As usual, to statistically model turbulent flows we consider the Navier-Stokes equation with a forcing term. Let $u^j(x,t)$ be the velocity field for an incompressible fluid, $\nabla\cdot u=0$. The Navier-Stockes equation with friction reads: \begin{eqnarray} \partial_t u^j + (u\cdot\nabla) u^j - \nu \nabla^2 u^j + \inv{\tau}\, u^j= -\nabla^jp + f^j \label{nseq} \end{eqnarray} with $p$ the pressure and $f(x,t)$ the external force such that $\nabla\cdot f=0$. The pressure is linked to the velocity by the incompressibility condition: $\nabla^2 p = -\nabla^i\nabla^j u^iu^j$. The friction term $\inv{\tau}\, u^j$ is introduced in order to mimick that in physical systems the infrared energy cascade terminates at the largest possible scale. The vorticity $\omega} \def\Om{\Omega$, with $\omega} \def\Om{\Omega=\epsilon_{ij}\partial_iu_j$, is transported by the fluid and satisfies: $\partial_t\omega} \def\Om{\Omega +(u\cdot\nabla)\omega} \def\Om{\Omega - \nu \nabla^2 \omega} \def\Om{\Omega +\omega} \def\Om{\Omega/\tau= F$ with $F=\epsilon_{ij}\partial_if_j$ \cite{cherk}. We choose the force to be gaussian, white-noice in time, with zero mean and two point function: $\vev{f^j(x,t)\ f^k(y,s)} = C^{jk}(x-y)\ \delta} \def\De{\Delta(t-s)$ where $C^{jk}(x)$, with $\nabla^j\, C^{jk}(x)=0$, is a smooth function varying on a scale $L_i$ and fastly decreasing at infinity. The correlation function of the vorticity forcing term is: $\vev{F(x,t)\ F(y,s)} = G(x-y)\ \delta} \def\De{\Delta(t-s) $ with $G=-\nabla^2 \widehat C$. We shall assume translation, rotation and parity invariance and expand $C^{jk}(x)$ as: $C^{jk}(x)= \overline \epsilon\, \delta} \def\De{\Delta^{ij} - \overline \epsilon_w\, (3r^2 \delta} \def\De{\Delta^{ij} - 2 x^ix^j)/8+ \cdots$ with $r^2=x^kx_k$. The scale $L_i$ represents the injection length. The inverse cascade takes place at distances $L_i\ll x \ll L_f\simeq \tau^{3/2}\, \overline \epsilon^{1/2}$. At finite viscocity, there are two ultraviolet characteristic lengths, the usual dissipative length $l_d \simeq \nu^{1/2}\, \overline \epsilon_w^{-1/6}$ and another friction length $l_f \simeq \nu^{1/2}\, \tau^{1/2}$ above which friction dominates over dissipation. The direct cascade takes place at scale $l_f\ll x \ll L_i$. We shall consider the inviscid limit at fixed friction, ie. $\nu\to 0$ at $\tau$ fixed. Note that $(l_f/\, l_d)^2 =\tau\, \overline \epsilon_w^{1/3}$ is large in the quoted experiments. The fundamental property of two dimensional turbulence recognized by Kraichnan \cite{kraich} and Batchelor \cite{batch} is that the energy cascades towards the large scales because it cannot be dissipated at small scales. In absence of friction the system does not reach a stationnary state. But friction provides a way for the energy to escape and for the system to reach a stationnary state. The fact that the energy is not dissipated at small scales translates into the vanishing of the averaged energy dissipation rate, ie. $\nu \vev{(\nabla u)^2} = 0$ in the inviscid limit, which means that there are no energy dissipative anomaly \cite{batch}, ie. \begin{eqnarray} \lim_{\nu\to 0} \nu (\nabla u)_{(x)}^2 = 0 \label{noano} \end{eqnarray} inside any correlation functions. However there could be enstrophy dissipative anomalies in the sense that $\nu (\nabla \omega} \def\Om{\Omega)^2$ does no vanish in the inviscid correlation functions: \begin{eqnarray} \widehat \epsilon_w(x) \equiv \lim_{\nu\to 0} \nu (\nabla \omega} \def\Om{\Omega)^2_{(x)} \label{ano} \end{eqnarray} Below we shall discuss possible consequences of the vanishing or not of $\widehat \epsilon_w$ in presence of friction. \medskip {\bf Two and three point velocity correlation functions.} In absence of energy dissipative anomaly, the mean energy density relaxes in the inviscid limit according to $\partial_t \vev{\frac{u^2}{2} } + \inv{\tau} \vev{u^2 } = \overline \epsilon $, showing that $\overline \epsilon$ is effectively the energy injection rate. It reaches a stationary limit with $\vev{u^2 }=\overline \epsilon\,\tau$. Similarly the enstrophy density evolves according to $ \partial_t \vev{\omega} \def\Om{\Omega^2} + 2\nu\vev{(\nabla\omega} \def\Om{\Omega)^2} + \frac{2}{\tau}\vev{\omega} \def\Om{\Omega^2} =2\overline \epsilon_w$ showing that $\overline \epsilon_w$ is the enstrophy injection rate. Let $\widehat \epsilon_w = \lim_{\nu\to 0} \nu\vev{(\nabla \omega} \def\Om{\Omega)^2}$ be the enstrophy dissipation rate, $\widehat \epsilon_w< \overline \epsilon_w$. In the stationary limit at finite friction the enstrophy density is finite and equal to: \begin{eqnarray} \vev{\omega} \def\Om{\Omega^2} = \tau \({ \overline \epsilon_w - \widehat \epsilon_w }\) \label{ombar} \end{eqnarray} Ie. the enstrophy density is equal to the difference of the enstrophy injection and the enstrophy dissipation rates times the friction relaxation time. The vorticity two point correlation function $\vev{\omega} \def\Om{\Omega(x)\omega} \def\Om{\Omega(0)}$ stay finite since it is bounded by $\vev{\omega} \def\Om{\Omega^2}$. Hence the vorticity correlation cannot diverge as $x\to 0$, and it cannot have a negative anomalous dimension. Scalings $\vev{\omega} \def\Om{\Omega(x)\omega} \def\Om{\Omega(0)}\sim r^{-\xi'}$ or $\sim (\log r)^{\xi'}$ with $\xi'>0$ are forbidden in presence of friction. At short distance one then may have: \begin{eqnarray} \vev{\omega} \def\Om{\Omega(x)\omega} \def\Om{\Omega(0)}_{\nu=0} = \overline \Om - (\tau \overline \epsilon_w) A\, (r/L)^{2\xi_2} + \cdots \ \label{omega} \end{eqnarray} with $\xi_2>0$. We shall denote the limiting value by $\overline \Om = \tau(\overline \epsilon_w - \eta_w)$. By eq({\ref{3point}) below, $\eta_w$ will be identified as the enstrophy transfer rate. The amplitude $A$, $\eta_w$ and $\widehat \epsilon_w$, as well as the anomalous exponant $\xi_2$, are functions of the dimensionless parameter $\tau^3\overline \epsilon_w$. Since $\nabla^2_x \vev{(\delta} \def\De{\Delta u)^2}=2 \vev{\omega} \def\Om{\Omega(x)\omega} \def\Om{\Omega(0)}$, finiteness of the vorticity two point function at coincidant points imply, $\vev{(\delta} \def\De{\Delta u)^2)} \simeq \frac{\overline \Om}{2}\, r^2$, but the subleading terms could have anomalous scalings: \begin{eqnarray} \vev{(\delta} \def\De{\Delta u)^2}= \frac{\overline \Om}{2}\, r^2 - (\tau\overline \epsilon_w )\frac{A}{2(\xi_2+1)^2} r^2\, (r/L)^{2\xi_2} + \cdots \label{2points} \end{eqnarray} The leading scaling is the normal Kraichnan scaling but the amplitude is different, ie. $\overline \Om$ instead of $\overline \epsilon_w^{2/3}$. There is a crossover at scale $r_c$, with $r_c^{2\xi_2}\sim(\overline \Om/\tau\overline \epsilon_w)$, above which the second anomalous contribution dominates. It is interesting to compare the occurence of both the smooth $r^2$ term in eq.(\ref{2points}) and the anomalous contribution $r^{2\xi_2}$ in eq.(\ref{omega}) with the passive scalar problem with friction \cite{cherk} in which anomalous zero modes occur in the passive scalar correlations only if the velocity flow is regular. The smooth $r^2$ term in $\vev{(\delta} \def\De{\Delta u)^2}$ does not contribute to the asymptotic large $k$ behavior of the energy spectrum which thus scales as $k^{-3-2\xi_2}$. Namely: \begin{eqnarray} E(k) \simeq -\frac{2^{2\xi_2+1}\Ga(1+\xi_2)}{\Ga(-\xi_2)}\, (\tau \overline \epsilon_w)A\, k^{-3}\, (kL)^{-2\xi_2} \label{spectrE} \end{eqnarray} as $k\to \infty$. The exponant $\xi_2$ is expected to be universal but not the amplitude $A$. As usual, stationarity of the two point correlation functions gives in the inviscid limit, for $x\not= 0$: \begin{eqnarray} \nabla_x^k \vev{ (\delta} \def\De{\Delta u^k)\, (\delta} \def\De{\Delta u)^2 } +2\vev{(\delta} \def\De{\Delta u)^2}/\tau &=& 2 (2\overline \epsilon - \widehat C(x) ) \simeq \overline \epsilon_w\, r^2 + \cdots \label{dissip} \end{eqnarray} It allows us to determine exactly the three point velocity inviscid structure functions: \begin{eqnarray} \vev{(\delta} \def\De{\Delta u)^3_{\|}}&=& \frac{\eta_w}{8}r^3 + \frac{3A\overline \epsilon_w}{4(\xi_2+1)^2(\xi_2+2)(\xi_2+3)} \, r^{3+2\xi_2} +\cdots \nonumber \\ \vev{(\delta} \def\De{\Delta u)_{\|}(\delta} \def\De{\Delta u)_{\bot}^2}&=& \frac{\eta_w}{8}r^3 + \frac{3(2\xi_2+3)A\overline \epsilon_w}{4(\xi_2+1)^2(\xi_2+2)(\xi_2+3)} \, r^{3+2\xi_2} +\cdots \label{3point} \end{eqnarray} {\bf Without dissipative anomaly at finite friction.} If there is no enstrophy dissipative anomalies in presence of friction, $\widehat \epsilon_w=0$ and hence $\eta_w=0$ assuming continuity of the correlations. The three point velocity structure (\ref{3point}) has an anomalous scaling at leading order. It behaves as $$\vev{(\delta} \def\De{\Delta u)^3} \sim r^{3+2\xi_2}$$ since the first term $\eta_wr^3$ vanishes. As a consequence the enstrophy transfer is not constant through the scales, ie. there is no enstrophy cascade. The anomalous scaling manifests itself only at a subleading order in the velocity two point function but the enstrophy structure function is anomalous: $$ \vev{(\delta} \def\De{\Delta \omega} \def\Om{\Omega)^2} = 2 (\tau \overline \epsilon_w)A(r/L)^{2\xi_2} $$ Moreover in absence of enstrophy dissipative anomaly the one point vorticity correlations are gaussian, \begin{eqnarray} \vev{\exp(s\omega} \def\Om{\Omega)} = \exp(s^2(\tau \overline \epsilon_w)/2) \label{omgauss} \end{eqnarray} This is a direct consequence of the stationarity of the vorticity correlations. It is not true for the higher point functions. It could provide a way to check whether the enstrophy dissipation vanishes or not in presence of friction. Below we shall present a perturbative argument in favor of the absence of enstrophy dissipative anomalies in presence of friction. In this scenario with $\widehat \epsilon_w=0$, the quasi symmetry of the probability distribution may be explained by the fact that the even order correlations are parametrically larger that the odd order correlations, as it is for the two and three point function, cf eqs.(\ref{2points},\ref{3point}) with $\eta_w=0$. Namely the even order correlation have leading normal but anomalous subleading contributions whereas the odd order correlations would have anomalous contributions at leading order which decrease faster at small scales. \medskip {\bf A scenario for the frictionless limit.} Experimentally, $(\tau^3\overline \epsilon_w)$ is large and we are thus interested in the frictionless limit. We shall discuss this limit assuming the absence of dissipative anomaly at $\tau$ finite. Demanding that the three point structure function reproduces the known frictionless exact result with $\vev{(\delta} \def\De{\Delta u)^3} \sim r^3$ as $\tau\to\infty$ imposes that the anomalous exponant $\xi_2$ vanishes and that the amplitude $A$ goes to one as $\tau \to \infty$. Moreover, demanding that the velocity structure function is finite as $\tau \to \infty$ requires: \begin{eqnarray} \xi_2 = \zeta_2\, (\tau^3\overline \epsilon_w)^{-1/3} + o((\tau^3\overline \epsilon_w)^{-1/3}) \quad {\rm and}\quad A=1 + c\,(\tau^3\overline \epsilon_w)^{-1/3} + o((\tau^3\overline \epsilon_w)^{-1/3}) \label{taularge} \end{eqnarray} with $\zeta_2$ and $c$ constant. Since $\xi_2$ is supposed to be universal but not the amplitude $A$, we expect $\zeta_2$ universal but not $c$. As a consequence, at fixed non zero distance $r$, the vorticity correlation behaves logarithmically as: \begin{eqnarray} \vev{\omega} \def\Om{\Omega(x)\omega} \def\Om{\Omega(0)}\simeq \overline \epsilon_w^{2/3}\, \[\, c - 2 \zeta_2 \log(r/L)\,\] + o((\tau^3\overline \epsilon_w)^{-1/3}) \label{omlarge} \end{eqnarray} Here we assume that the terms represented by the dots in eq.(\ref{omega}) which are subleading, and thus negligeable, at finite friction do not become predominant in the frictionless limit. But this is the simplest scenario. The vorticity structure function, estimated as $\vev{(\delta} \def\De{\Delta \omega} \def\Om{\Omega)^2} = 2\tau \overline \epsilon_w -2 \overline \epsilon_w^{2/3} \,( c - 2 \zeta_2 \log(r/L))$, is then dominated by the constant term since the enstrophy density diverges as $\tau \to \infty$. This may explain the experimental fact that vorticity correlations are almost constant in the inertial range. This scenario is compatible with experimental data since experimentally $ \vev{(\delta} \def\De{\Delta \omega} \def\Om{\Omega)^2}\simeq 10\, s^{-2}$ whereas $2\tau \overline \epsilon_w \simeq 20\, s^{-2}$ with an estimated error of $20\,{}^0/_0$ to $30\,{}^0/_0$. The velocity structure function will then be: $$\vev{(\delta} \def\De{\Delta u)^2} \simeq \overline \epsilon_w^{2/3}\,r^2\, \[ c/2 + \zeta_2 - \zeta_2 \log(r/L) + \cdots\] $$ The smooth $r^2$ is not universal but the logarithmic correction is (since $\zeta_2$ is expected to be universal). In the same way the energy spectrum (\ref{spectrE}) $E(k)$ becomes proportional to $k^{-3}$ with: \begin{eqnarray} E(k)\simeq 2\zeta_2\, \overline \epsilon_w^{2/3}\, k^{-3} \label{krachn} \end{eqnarray} with Kraichnain's constant $C_K=2\zeta_2$ which is universal if the anomalous dimensions at $\tau$ finite are universal. Note that the smooth but non universal $r^2$ terms in the velocity correlation does not contribute to the energy spectrum at large momenta. Of course, there could be logarithmic correction to the spectrum if $\zeta_2=0$. \medskip {\bf With dissipative anomaly at finite friction.} If the enstrophy dissipation is non zero in presence of friction, the enstrophy density is finite but smaller than the injected enstrophy density: $\vev{\omega} \def\Om{\Omega^2} < \tau \overline \epsilon_w$. The three point velocity structure function then scales as $\vev{(\delta} \def\De{\Delta u)^3} \sim \eta_w r^3$ which allows us to identify $\eta_w$ as the enstrophy transfer rate, ie. there is enstrophy cascade with transfer rate $\eta_w$. The quasi symmetry of the probability distribution function would hold only if the entrophy transfer rate is much smaller that the entrophy injection rate $\eta_w\ll \overline \epsilon_w$. There are many scenario for the frictionless limit depending on how $\eta_w/\overline \epsilon_w$, which is a function of $\tau^3\overline \epsilon_w$, behaves as $\tau\to \infty$. \medskip {\bf Neglecting odd order correlations.} We now discuss what would be the consequences of neglecting the odd order correlations in front of the even order ones. This is suggested by the quasi symmetry of the experimentally measured probability distribution. Consider the equations encoding the stationarity of an even number of velocity differences, ie. $\partial_t\vev{(\delta} \def\De{\Delta u^{i_1})\cdots (\delta} \def\De{\Delta u^{i_{2n}})}=0$. Both the terms arising from the advection $(u\cdot\nabla) u$ or from the pressure $\nabla p$ involve correlation functions with an odd number of velocity insertions. Assuming that the odd order velocity correlations are much smaller that the even order ones, these terms are much smaller and negligeable compared to the remaining terms which are those arising from the friction and from the forcing and which involve an even number of velocity insertions. Hence at the leading order the stationarity condition of the even order correlations reads: \begin{eqnarray} \frac{2n}{\tau} \vev{(\delta} \def\De{\Delta u^{i_1})\cdots (\delta} \def\De{\Delta u^{i_{2n}})}^{(0)}= 2 \sum_{p<q} (\delta} \def\De{\Delta C^{i_pi_q}_{(x)})\, \vev{\cdots \widehat {(\delta} \def\De{\Delta u^{i_p})}\cdots \widehat {(\delta} \def\De{\Delta u^{i_{q}})} \cdots}^{(0)} \label{statn} \end{eqnarray} with $\delta} \def\De{\Delta C^{ij}(x) = C^{ij}(0)-C^{ij}(x)$. The overhatted quantities are omitted in the correlation functions. In eqs.(\ref{statn}) we have used the fact that there is no energy dissipative anomalies. The solution of eqs.(\ref{statn}) is provided by a gaussian statistics with zero mean and two-point function $\vev{ (\delta} \def\De{\Delta u^i)(\delta} \def\De{\Delta u^j) }^{(0)} \simeq \frac{\tau\, \overline \epsilon_w}{8}( 3r^2 \delta} \def\De{\Delta^{ij} - 2 x^ix^j)$. Hence, in this approximation the velocity structure functions scale as: \begin{eqnarray} F^{(0)}_{2n}\equiv \vev{ (\delta} \def\De{\Delta u)^{2n} }^{(0)} &\simeq& {\rm const.}\ (\tau \overline \epsilon_w r^2)^n \label{ansatz} \end{eqnarray} This is compatible with eq.(\ref{2points}) since $\overline \Om=\tau \overline \epsilon_w$ if $\eta_w \ll \overline \epsilon_w$. To be consistent it also has to solve the stationarity equations for the odd order correlations which at this order read: \begin{eqnarray} \nabla_{x_1}^k \vev{ u^k_{(x_1)}u^{i_1}_{(x_1)}\, u^{i_2}_{(x_2)} \cdots u^{i_{2n-1}}_{(x_{2n-1})} }^{(0)} + \nabla_{x_1}^{i_1} \vev{ p_{(x_1)}\, u^{i_2}_{(x_2)} \cdots u^{i_{2n-1}}_{(x_{2n-1})}}^{(0)} + {\rm perm.} = O(r^{2n}) \label{press} \end{eqnarray} with $\nabla^2 p = - \nabla^j\nabla^k u^ju^k$. Since the statistics is gaussian it is enough to check them for $n=2$. The gradiant of the pressure has scaling dimension one. Then the l.h.s. up to $O(r^4)$ is a rank three symmetric $O(2)$ polynomial tensor, transverse along all coordinates, invariant by translation and of scaling dimension three. There is no such tensor and thus the l.h.s. is $O(r^4)$ for $n=2$. In other words, the $n$-point velocity correlations computed using the gaussian statistics are zero modes. In view of the exact formula eqs.(\ref{2points},\ref{3point}), this approximation would be better for the transverse velocity correlations. Note also that using this approximation to compute the one point vorticity functions $\vev{\omega} \def\Om{\Omega^n}$ gives the exact result (\ref{omgauss}) if there is no enstrophy dissipative anomaly. Of course there would be deviations to this approximation: $$F_{n} \simeq F^{(0)}_{n} + \delta} \def\De{\Delta F_{n}$$ with corrections $\delta} \def\De{\Delta F_{n}$ including both contributions with normal scaling or with possible anomalous scaling as in eq.(\ref{2points},\ref{3point}). In the scenario without enstrophy dissipation, these anomalous contributions will be subleading in the even order correlation but dominant at small scales in the odd order correlations. On contrary if the enstrophy transfer is not vanishing these corrections include contributions with normal scaling, since $\delta} \def\De{\Delta F_2\simeq \eta_w\tau r^2$ and $\delta} \def\De{\Delta F_3 \simeq \eta_w r^3$ as follows from eqs.(\ref{2points},\ref{3point}). Stationnarity of the three point functions leads then to an equation for the four point correlations of the form $\nabla^\pi F_4 +\inv{\tau} F_3=0$ with $\nabla^\pi$ the transverse gradiant. Using $F_3 \simeq \eta_w r^3$ and naive scaling, ie. omitting possible zero mode contributions, gives a correction $\delta} \def\De{\Delta F_4 \simeq \frac{\eta_w}{\tau} r^4$ to the first order gaussian contribution $F_4^{(0)}$ which satisfies $\nabla^\pi F_4^{(0)}=0$. Similarly, stationarity of the four point functions now give equations of the form $\nabla^\pi F_5 +\inv{\tau} \delta} \def\De{\Delta F_4 \simeq \eta_w \tau\overline \epsilon_w r^4$. For $\tau^3\overline \epsilon_w \gg 1$ as in \cite{tabis}, one may neglect $\inv{\tau} \delta} \def\De{\Delta F_4$ in front of $\eta_w \tau\overline \epsilon_w r^4$, so that $\delta} \def\De{\Delta F_5 \simeq \eta_w \tau \overline \epsilon_w r^5$. Recursively, naive dimensional analysis apply to stationarity equations with $\tau^3 \overline \epsilon_w \gg 1$ yield to first order in $\eta_w$: $\delta} \def\De{\Delta F_{n+2} \simeq \eta_w\, \tau^{1-n}\, (\tau^3 \overline \epsilon_w)^k\, r^{n+2}$ with $k$ the integer part of $n/3$. Of course $\delta} \def\De{\Delta F_{2n}\ll F^{(0)}_{2n}$ are small for $\tau^3 \overline \epsilon_w \gg 1 \geq (\eta_w/\overline \epsilon_w)$. But zero modes of the stationarity equations which have been neglected in this dimensional analysis may give additional contributions to $\delta} \def\De{\Delta F_n$. These zero modes will be the leading terms for the odd order correlations if the enstrophy dissipation and thus the enstrophy transfer rate vanish. \medskip {\bf MSR formalism.} We now discuss how these scenario could fit in a field theory approach to the problem. The corrections (\ref{taularge}) of the anomalous exponant as a function of the friction coefficient have a natural interpretation in the MSR formalism \cite{msr}. The MSR formalism provides a way to compute the inviscid vorticity correlations using a path integral with action \begin{eqnarray} S= i\int dt dx\, \varphi_{(x,t)}(\partial_t\omega} \def\Om{\Omega + u\cdot \nabla \omega} \def\Om{\Omega + \inv{\tau}\omega} \def\Om{\Omega)_{(x,t)} + {1 \over 2} \int dt dx dy\, \varphi_{(x,t)}G_{(x-y)}\varphi_{(y,t)} \label{action} \end{eqnarray} Note that this action describes the inviscid limit $\nu=0$. Let us first define $\omega} \def\Om{\Omega(x,t)=\overline \epsilon_w^{1/3} \widetilde \omega} \def\Om{\Omega(t\overline \epsilon_w^{1/3},x)$ and $\varphi(x,t)=\overline \epsilon_w^{-1/3} \widetilde \varphi(t\overline \epsilon_w^{1/3},x)$. This corresponds to select configurations for which the dominant time scales and amplitudes are $\overline \epsilon_w^{-1/3}$. It allows us to present the theory as a perturbation of the frictionless action $S_0$: \begin{eqnarray} S = S_0 + ig^{-1/3} \int dt dx\,\widetilde \varphi_{(x,t)}\widetilde \omega} \def\Om{\Omega_{(x,t)} \label{pertfric} \end{eqnarray} with $g=\tau^3\overline \epsilon_w$ as dimensionless expansion parameter. This shows that the pertubation around the frictionless theory is controled by $(\tau^3\overline \epsilon_w)^{-1/3}$ as in eq.(\ref{omlarge}). Of course implementing this perturbation remains a challenge as the unperturbed theory is unkown. Alternatively let us define $\omega} \def\Om{\Omega(x,t)=(\tau \overline \epsilon_w)^{1/2} \omega} \def\Om{\Omega'(t/\tau,x)$ and $\varphi(x,t)=(\tau \overline \epsilon_w)^{-1/2} \varphi'(t/\tau,x)$. This corresponds to select configurations with time scale $\tau$. The theory is then formulated as a perturbation of the gaussian action by the advection term with $g^{1/2}=(\tau^3\overline \epsilon_w)^{1/2}$ as coupling constant: \begin{eqnarray} S= S_{{\rm gauss}} + i g^{1/2}\int dt dx\, \varphi'_{(x,t)} (u'_{(x,t)}\cdot \nabla )\omega} \def\Om{\Omega'_{(x,t)} \label{pertadv} \end{eqnarray} where $S_{{\rm gauss}}$ is the gaussian action obtained by neglecting the advection term. This is the statistics we found by neglecting odd order correlation functions. In the unperturbed theory $\vev{\omega} \def\Om{\Omega^2}=\tau \overline \epsilon_w$ and there is no anomaly. Perturbation theory is then computable. We checked that there is no occurence of the dissipative anomaly in first order (one loop) in the perturbation theory. This is linked to the fact that the friction regularizes the theory in the inviscid limit such that there is no divergence in the Feymann amplitudes, at least in the first order we checked. This is expected to be true to all orders. In other words, there is no pertubative contribution to the enstrophy transfer rate, ie $\widehat \epsilon_w=\eta_w\simeq 0$ in perturbation theory. Non perturbative contributions may arise due to instanton contributions to the path integral. Let us for exemple define $\omega} \def\Om{\Omega(x,t)=\tau^{-1} \widehat \omega} \def\Om{\Omega(t/\tau,x)$ and $\varphi(x,t)=(\tau^2 \overline \epsilon_w)^{-1} \widehat\varphi(t/\tau,x)$. It allows us to write the action as $S=(\tau^3\overline \epsilon_w)^{-1} \widehat S$ with $\widehat S$ dimensionless. In the saddle point approximation, instantons could then potentially give a non perturbative contribution to the enstrophy transfer rate with $\eta_w \sim ({\rm const.})\tau^{-3} \exp(-{\rm const.}/ (\tau^3\overline \epsilon_w))$. \bigskip {\bf Acknowledgements:} We thank P. Tabeling for motivating discussions on their experiments and M. Bauer and K. Gawedzki for numerous friendly conversations. This research is supported in part by the CNRS, by the CEA and the European TMR contract ERBFMRXCT960012.
\section{Figures} \begin{figure} \caption{The wavefunction of an incommensurate ground state having winding number 89/144 at fixed $\tilde{\hbar}=0.2$ for different values of $K$. The wavefunction $\langle|\Psi|^2\rangle$ means the probability of finding the particles at $X$ (mod $2\pi$). The curves are for $K=0.1, 1, 2$, and $5$, respectively. The wavefunction becomes localized at the lower part of the potential as $K$ changes to 2 and 5. } \end{figure} \begin{figure} \caption{The quantum ``disorder parameter'' $P_t$, i.e. the probability of finding particles on the top of the external potential versus $K$ for a fixed quantum fluctuation $\tilde{\hbar}=0.2$. (a) $P_t$ for different temperature at a fixed winding number 13/21. The temperature changes from $0.2/30, 0.2/120$, and $0.2/480$, respectively. (b) $P_t$ for different winding numbers: 13/21, 34/55, and 89/144, at fixed temperature $T=0.2/120$, which can be effectively regarded as zero. } \end{figure} \begin{figure} \caption{ $\Delta$ versus $K$. Different symbols represent different effective Planck's constant, which are $\tilde{\hbar}= 0.01$, and 0.1, respectively. We draw the classical disorder parameter defined by Coppersmith and Fisher [17] in the inset for comparison. The winding number for all calculations given in this figure is 34/55. } \end{figure}
\section{Introduction} Strong interactions at low energies can be described by an effective field theory (EFT) of pions and nucleons with the nucleons treated nonrelativistically.~\cite{all} Chiral symmetry prescribes how pions interact. All other interactions are parameterized by the most general local Lagrangian consistent with the symmetries of quantum chromodynamics (QCD). The coefficients of these general terms can either be predicted directly from QCD, which may be possible in future lattice calculations, or fit from existing data. I will focus on nucleon-nucleon (NN) scattering in the following. The data here is quite good and should lead to an excellent determination of these constants. However, since nuclear forces produce bound states, a nonperturbative treatment of the EFT is required, which adds a new twist to this procedure.~\cite{all} Any EFT will break down at a scale associated with the underlying physics not taken into account in the original Lagrangian. For nuclear EFTs, an optimist may believe this will occur when the center of mass momentum of the nucleons excite either the exchange of the next heaviest meson after the pion, $p\sim m_\rho$, or the nucleon-$\Delta$ mass difference $p=\sqrt{M (M_\Delta-M)}\sim 525$~MeV. In practice, the breakdown scale seems to be much lower, $p\sim 300$~MeV. This could be due to anything from quark substructure to subtleties in the fitting. Furthermore, two different organizational schemes for calculations in the EFT seem to imply different physical reasons for the breakdown. This would be unfortunate, since alternative ways to organize the corrections should have no effect on the physics. By using some models, I will show what actually does set this scale and then use this to make connections to the real world. Extending the applicability of the EFT to the scale associated with the $\Delta$ seems like a small achievement, but could make calculations of nuclear matter converge much faster. In that case, the momentum of interest is on the order of the Fermi momentum at saturation $p_F\sim 280$~MeV, which currently is too close to the breakdown scale for an EFT analysis to be useful. \section{Removing vs. Weakening the Pion}\label{rm} Assume, for now, that interactions between two nucleons are exactly given by potential exchange of both a long- and short-ranged meson with masses $m_\pi=140$~MeV and $m_\rho=770$~MeV respectively, \begin{equation} V_{\rm exact}(r) = - \alpha_\pi \frac{e^{-m_\pi r}}{r} - \alpha_\rho \frac{e^{-m_\rho r}}{r}\, . \label{yuks} \end{equation} The pion coupling is taken to be $\alpha_\pi=g_A^2 m_\pi^2/(16 \pi f_\pi^2)\simeq0.075$, as in the real world, and the rho coupling $\alpha_\rho=1.05$ is tuned to give a large scattering length $a=-23.4$~fm, as observed in data for the ${}^1\!S_0$ partial wave of NN scattering. The following discussion does not depend on the value of the potential at the origin, so I will take $V_{\rm exact}$ to be zero there for simplicity. \begin{figure} \begin{center} \leavevmode \epsfxsize=2.3in \epsffile[84 621 201 725]{cut.ps} \end{center} \caption{\label{cut}The analytic structure of an observable for the two-Yukawa model of Eq.~(\protect\ref{yuks}) in the complex momentum plane.} \end{figure} This potential generates observables with analytic structure as shown in Fig.~\ref{cut}, specified by a cut at both $m_\pi/2$ and $m_\rho/2$. The existence of these cuts indicates that a low-momentum expansion of an observable will not converge for momenta above $m_\pi/2$. An EFT can only go beyond this point if it either exactly reproduces or sufficiently approximates the analytic structure of the pion. This can be illustrated by focusing on the most general Lagrangian without pions (I will restrict myself to the ${}^1\!S_0$ channel below) \begin{equation} {\cal L}_{\rm EFT} = N^\dagger \left[ i \partial_t + \frac{\nabla^2}{2M} \right] N - C_0 (N^\dagger N)^2 + \frac12 C_2 \left[ (N^\dagger \nabla N)^2 + h.c. \right] + \ldots \, , \label{Leff} \end{equation} which corresponds to a nonrelativistic potential between two nucleons of \begin{equation} V_{\rm EFT}(p,p') = C_0 + \frac12 C_2 \left( p^2 + p'{}^2\right) + \ldots \, , \label{Veff} \end{equation} where $p$ and $p'$ are the center of mass momentum of the incoming and outgoing nucleons. Since these interactions depend on powers of momentum, a naive calculation of the NN scattering phase shifts from ${\cal L}_{\rm eff}$ leads to ultraviolet divergences. This just reflects our ignorance of the high-energy behavior of this theory. Two possible procedures to regulate these divergences are: i) adding a cutoff to all integrals to eliminate or suppress the high momentum states and ii) using dimensional regularization, which amounts to throwing out any integrals that diverge like powers of momentum. The second procedure may seem a bit counterintuitive at first, but is simpler to implement since it does not break any symmetries of the underlying theory and therefore does not require additional counterterms. Perturbative calculations in chiral perturbation theory have confirmed the equivalence of these regularization procedures. Nonperturbative calculations, on the other hand, show that a naive implementation of dimensional regularization is incomplete. A generalization to account for integrals that linearly diverge is required. This regularization is called power divergence subtraction~\cite{PDS} (PDS). After regulating the theory, the EFT can be organized to give predictions order-by-order in powers of an expansion parameter. Since the low-energy constants $C_{2n}$ are not determined by the symmetries of QCD, they must be fit to data. In the process, they will typically pick up the next scale $\Lambda$ of physics not explicitly accounted for in the EFT, leading to $C_{2n} p^{2n} \sim (4\pi/\Lambda^2) (p/\Lambda)^{2n}$. Therefore interactions with more derivatives are suppressed by additional powers of $p/\Lambda$. Power counting in this parameter allows Eq.~(\ref{Leff}) to be truncated to the first few terms, with more accurate results obtained by keeping more terms in the momentum expansion. There are two prevalent ways of implementing power counting in nonperturbative effective field theories. The first is to count powers of $\Lambda$ in the potential $V_{\rm EFT}$ and then solve the Lippmann-Schwinger equation. Such a procedure sums an infinite number of Feynman diagrams to generate the nonperturbative amplitude ${\cal A}$, related to the phase shift $\delta$ by \begin{equation} {\cal A} = \frac{4\pi/M}{p\cot\delta-ip} \, . \end{equation} This method of power counting is called $\Lambda$-counting~\cite{lepage} and is in the spirit of the Hamiltonian theory of Wilson, first applied to nuclear EFTs by Weinberg.~\cite{Weinberg} An alternative is to expand ${\cal A}$ in powers of $Q\equiv (p,m_\pi,1/a)$, motivated by chiral perturbation theory, which expands in $p$ and $m_\pi$. The nonperturbative nature of the interactions identifies $1/a$ as a small parameter as well, justifying its inclusion in the expansion. This method of power counting is called $Q$-counting.~\cite{PDS,cohen} For a theory without pions, such as Eq.~(\ref{Leff}), the underlying scale $\Lambda$ is just $m_\pi$. Both power counting procedures work as long as momenta are less than this scale, reproducing the effective range expansion \begin{equation} p\cot\delta = - \frac1{a} + \frac12 p^2 \sum_{n=0}^\infty r_n \frac{p^{2n}}{\Lambda^{2n}} \, . \label{ERe} \end{equation} The effective ranges, which are determined from NN scattering data,~\cite{NNdata} also exhibit this scale $r_n\sim 1/\Lambda\sim 1/m_\pi$. To extend the EFT to higher momenta, pions need to be explicitly added to the effective Lagrangian. This hopefully will bring $\Lambda$ up to the QCD scale of $m_\rho$ or possibly even 1~GeV. The inclusion of pions nonperturbatively is different in the two power counting procedures. Every iteration of one-pion exchange contributes to the pion cut in Fig.~\ref{cut}. In general, $n$ iterations are proportional to% \footnote{This can be seen by studying the exact solution of an exponential potential,~\cite{us2} which has the same cut structure as a Yukawa or by looking at the analytic expressions of higher Yukawa loops, which exhibit the cut structure through Polylogarithms.~\cite{binger}} \begin{equation} \mbox{$n$-Pion Exchange} \propto \left[ \frac{m_\pi}{\Lambda_{{\rm NN}}} \frac{m_\pi^2}{4 p^2} \ln \! \left( 1 + \frac{4 p^2}{m_\pi^2} \right) \right]^n\, , \label{ope} \end{equation} where $\Lambda_{{\rm NN}}\simeq300$~MeV is just another way of expressing the strength of the interaction via the relation $\alpha_\pi=m_\pi^2/(M\Lambda_{{\rm NN}})$. $\Lambda$-counting includes this long-range part of one-pion exchange fully at leading order, removing the effect of the pion cut. $Q$-counting pushes one-pion exchange to subleading order and iterated one-pion exchange to even higher orders. Therefore, each order in the $Q$-counting expansion weakens the pion cut by an additional power of $m_\pi/\Lambda_{{\rm NN}}$, as seen in Eq.~(\ref{ope}). The topic of the next section is whether this simpler implementation of power counting, also called perturbative pions, can be as successful as $\Lambda$-counting. \section{New Fitting Procedures}\label{fits} The NN EFT is expected to work best at low-momentum. So a natural procedure to fix the low-energy constants $C_{2n}$ is to calculate an observable, like $p\cot\delta$, and fit near threshold. However, there are some complications originating from the nonperturbative nature of the problem. These must be cleared up before being able to make a statement about the breakdown scale. \begin{figure} \begin{center} \leavevmode \hbox{ \hspace{-1.25cm} \epsfxsize=2.5in \epsffile{ExpModel.eps} \hspace{.25cm} \epsfxsize=2.5in \epsffile{ExpMER.eps} } \end{center} \caption{\label{expmod} The error in $p\cot\delta$ for a pion coupling three times stronger than the real world. The left plot is from a normal fit to the low-energy constants both without (solid) and with (dashed) explicit pions. The right plot uses the modified effective range expansion to fit.} \end{figure} \subsection{$\Lambda$-counting and the Modified Effective Range Expansion} Most calculations that implement $\Lambda$-counting require a numerical treatment (although there are exceptions~\cite{cohen}). A nice way to show the systematic improvement expected of an EFT in this case is to use error plots~\cite{lepage,us1,us2} as in Fig.~\ref{expmod}. As more and more constants are added to the effective potential Eq.~(\ref{Veff}), the error in $p\cot\delta$ improves systematically by powers of $p^2$, reflected in the improving slopes. The point at which all errors are of the same order is where the theory is expected to break down.~\cite{us1} The solid lines of the left plot of Fig.~\ref{expmod} show the case without explicit pions. The breakdown point, or radius of convergence, of the EFT is around $m_\pi/2$, as expected from the discussion in Section~\ref{rm}. Adding pions to the short-ranged Lagrangian Eq.~(\ref{Leff}) as dictated by chiral symmetry should improve the radius of convergence.~\cite{bira} However, one needs to be careful when implementing this with cutoff regularization. If the pion coupling were three times stronger, for example, adding it to the Lagrangian would {\it not} improve the breakdown scale.~\cite{us2} This is illustrated by the dashed lines in the left plot of Fig.~\ref{expmod}. Chiral symmetry helps in that the coupling is small, but an EFT should work regardless of the strength of the potential being added. What went wrong? \begin{figure} \begin{center} \leavevmode \epsfxsize=5in \hbox{ \hspace{-.5in} \epsffile[121 578 538 674]{pions.ps} } \end{center} \caption{\label{pions}Symbolic definition of quantities in the modified effective range function in Eq.~(\protect\ref{MERF}), which sums all the pion contributions.} \end{figure} Note that the effective ranges $r_n$ of Eq.~(\ref{ERe}), which are used to fix the values of the $C_{2n}$, exhibit the scale $m_\pi$. If this scale were completely removed from the fitting procedure, then the constants should take on the proper scale $m_\rho$. One way to do this is to construct a function similar to $p\cot\delta$ which is analytic all the way to the rho cut in Fig.~\ref{cut}. In proton-proton scattering analyses, such a procedure is used to remove the threshold cuts coming from photon exchange. Called the modified effective range expansion,~\cite{modER} it can be used here to remove the effect of pion exchange by defining \begin{equation} K(p) \equiv \frac{p\cot(\delta-\delta_\pi)}{|{\cal F}_\pi(p)|^2} + \mbox{Re}\, \left[ \frac{f_\pi'(p,0)}{{\cal F}_\pi(p)} \right] \equiv -\frac1{\wta} + \frac12 p^2 \sum_{n=0}^\infty \widetilde{r}_n \frac{p^{2n}}{\widetilde{\Lambda}^{2n}} \ . \label{MERF} \end{equation} All possible insertions of one-pion exchange are taken into account in this formula, as seen pictorially in Fig.~\ref{pions}. For the model potential Eq.~(\ref{yuks}), this function is analytic up to $m_\rho/2$ and has modified effective ranges $\widetilde{r}_n\sim 1/\widetilde{\Lambda} \sim 1/m_\rho$. Fitting the $C_{2n}$'s using the function $K(p)$ leads to much better results, shown in the right plot of Fig.~\ref{expmod}. The radius of convergence is now on the order of $m_\rho$ as expected. Using a cutoff to regulate the theory mixes long- and short-distance physics, which is what led to the premature breakdown.~\cite{us2} The modified effective range expansion removes the long-distance physics from the observable, allowing a clean fit of the low-energy constants to the short-distance physics. In principle, this can also be achieved by the addition of more counterterms, but the modified effective range expansion is a simpler alternative. Luckily, one-pion exchange in the real world is weak enough that an improper treatment of this malady does not make a difference.~\cite{us2} However, this might not be the case for two-pion exchange as will be discussed in Section~\ref{2pi}. \begin{figure} \begin{center} \leavevmode \epsfxsize=4in \epsfysize=3.7in \epsffile{WeinKSW.eps} \end{center} \caption{\label{qcount}Comparison of the two power counting procedures for the two-Yukawa model Eq.~(\protect\ref{yuks}) using a fit at low momentum.} \end{figure} \subsection{$Q$-counting and a Global Fit} A low-momentum fit to data using $Q$-counting and perturbative pions also breaks down much earlier than expected.~\cite{us2} In Fig.~\ref{qcount}, the leading order (LO) and next-to-leading order (NLO) calculations using $Q$-counting are compared to those using $\Lambda$-counting. The $\Lambda$-counting result has a radius of convergence around $m_\rho$ after the low-energy constants are matched to the modified effective range expansion. Implementing the modified effective range expansion is inconsistent with the idea of perturbative pions, and so the $Q$-counting result breaks down around $m_\pi$. However, it is important that the low-energy constants only pick up the (unknown) short-distance physics, and matching to the modified effective range expansion is the most convenient way to do this. Using PDS, this would lead to the expressions~\cite{kap} \begin{equation} C_0 = \frac{4\pi/M}{-\mu+1/\widetilde{a}} \, , \qquad C_2 = \frac{MC_0^2}{4\pi} \frac12 \widetilde{r_0} \, , \qquad\ldots\, , \label{modPDS} \end{equation} which only depend on the short-distance physics.% \footnote{Although the renormalization scale $\mu$ appears in these low-energy constants, all observables are independent of it.} One important modification to this would be to match the pole of the amplitude ${\cal A}$ exactly~\cite{mehstew} \begin{equation} C_0 = \frac{4\pi/M}{-\mu+1/a} \, . \label{c0} \end{equation} Otherwise, terms attempting to correct the position of the pole would enter as double poles at higher orders messing up the power counting. This modification requires perturbative corrections to $C_0$ that compensate the pion contributions at each order. However, this still leaves the other $C_{2n}$'s dependent on the modified effective ranges $\widetilde{r}_n$, which cannot be obtained using perturbative pions. \begin{figure} \begin{center} \leavevmode \epsfxsize=3in \epsffile{Fit.eps} \end{center} \caption{\label{toy}$Q$-counting expansion of the two delta-shell model. The exact expression (solid) coincides nicely with the EFT (dashed) when a global fit is minimized with respect to the modified effective ranges. The dashed line at LO is on top of the solid line.} \end{figure} A method must therefore be devised to carry out the matching of Eq.~(\ref{modPDS}). It is reasonable to assume that when the modified effective ranges are set to the correct values, the error of the EFT will be less than when the ranges are at some random values. This can be checked by using a model. Replacing the Yukawas in Eq.~(\ref{yuks}) by delta-shells \begin{equation} - \alpha_i \; \frac{e^{-m_i r}}{r} \to - g_i \; \delta\!\left(r-\frac1{m_i} \right) \, , \end{equation} an analytic expression for the exact answer can be obtained.~\cite{kap} Expanding the exact amplitude in powers of $Q$, it can be compared to the blind EFT fit of the low-energy constants to data using $Q$-counting. The results coincide best when the modified effective ranges are varied to globally minimize the $\chi^2$ for data in a representative momentum range like $[m_\pi,m_\rho]$. The results are shown up to next-to-next-to-leading order (NNLO) in Fig.~\ref{toy}. They suggest that for $Q$-counting, a constrained global fit of the EFT to data to fix the low-energy constants is better than a low-momentum fit. The constraint is that the modified effective ranges, not the low-energy constants themselves, are varied to achieve the best fit. One thing to notice about Fig.~\ref{toy} is that the errors do not show the systematic improvement in slope characteristic of the earlier error plots. Although the scattering length is fit exactly to reproduce the pole in the amplitude Eq.~(\ref{c0}), the other terms in the effective range expansion have corrections at each order in the $Q$-counting. A comparison of Fig.~\ref{toy} with the right plot of Fig.~\ref{expmod} shows that accuracy is sacrificed near threshold to obtain the global fit needed for this power counting. In addition, the ability to read the EFT radius of convergence from the error plot is gone. Another way to estimate the breakdown scale is therefore needed. The calculated amplitude ${\cal A}$ starts at order $Q^{-1}$ and has a contribution ${\cal A}_n$ at each order $Q^n$ in the expansion. Up to this point, I have extracted the $Q$-counting phase shift by expanding the expression~\cite{PDS} \begin{eqnarray} \delta &=& \frac1{2i}\; \ln\!\left[ 1 + i \frac{Mp}{2\pi} \left( {\cal A}_{-1} + {\cal A}_0 + \ldots \right) \right] \label{exp1} \\ &\to& \frac1{2i}\; \ln\!\left( 1 + i \frac{Mp}{2\pi} {\cal A}_{-1} \right) + \frac{Mp}{4\pi} \left( \frac{{\cal A}_0}{1+i \frac{Mp}{2\pi} {\cal A}_{-1}} \right) + \ldots \, , \label{exp2} \end{eqnarray} but I could also have kept the full expression Eq.~(\ref{exp1}). These two expressions are equivalent up to terms that are higher order in the $Q$ expansion. Those extra terms, though, are an estimate of the corrections to the actual result and depend on the radius of convergence. The breakdown for $Q$-counting can be identified as the point at which the two expressions Eqs.~(\ref{exp1}) and (\ref{exp2}) diverge from each other. Doing this exercise for the two-Yukawa model results in Fig.~\ref{2yuk}. There it is easy to see the actual breakdown is somewhere around $m_\rho$ as expected. \begin{figure} \begin{center} \leavevmode \epsfxsize=3.5in \epsffile{2Yuk.eps} \end{center} \caption{\label{2yuk}Determination of the $Q$-counting breakdown scale for the two-Yukawa model.} \end{figure} \section{Data and the Role of Two-Pion Exchange}\label{2pi} Now that the two power counting methods are well understood and give the proper radius of convergence for a model, it is appropriate to apply them to actual NN-scattering data. Using the Nijmegen partial wave analysis,~\cite{NNdata} the radius of convergence for each of the two different power counting schemes is shown in Fig.~\ref{data}. The left plot shows that, for $\Lambda$-counting with one-pion exchange included as in the previous section, the breakdown is around $p\sim 300$~MeV. The right plot shows a similar scale for $Q$-counting with perturbative pions. This scale is less than expected and seems to be set by physics other than the $\rho$ or $\Delta$. Early calculations in $Q$-counting predicted a breakdown at $\Lambda_{{\rm NN}}$, which would coincide with the observed scale. However, recall that in the previous section I showed that $Q$-counting broke down around $m_\rho$ in the two-Yukawa model. That calculation had identical content with respect to one-pion exchange, indicating that the $300$~MeV comes from different physics. A closer study of $\Lambda$-counting determines that two-pion exchange should be included at NLO.~\cite{bira} \begin{figure} \begin{center} \vspace*{-.3cm} \leavevmode \hbox{ \hspace{-1.25cm} \epsfxsize=2.5in \epsffile{Dat2.eps} \hspace{.25cm} \epsfxsize=2.5in \epsffile{Dat.eps} } \end{center} \caption{\label{data} Results for ${}^1\!S_0$ NN-scattering data. The breakdown scale is determined by the left plot for $\Lambda$-counting and the right plot for $Q$-counting.} \end{figure} The fact that the actual data behaves so much differently from the two-Yukawa model indicates the model was not sophisticated enough. Furthermore, no reasonable choices for the parameters of this model can give as large an effective range $r_0=2.63$~fm as seen in the ${}^1\!S_0$ channel.~\cite{us2} Until now I have assumed that $m_\pi$ and $m_\rho$ are the underlying scales, but nuclear physics wisdom tells us there are actually two short-range Yukawas \begin{equation} V(r) = - \alpha_\pi \frac{e^{-m_\pi r}}{r} - \alpha_\sigma \frac{e^{-m_\sigma r}}{r} + \alpha_\rho \frac{e^{-m_\rho r}}{r} \, , \label{3yuks} \end{equation} given by an attractive scalar (of mass between $500$ and $600$~MeV) and repulsive vector particle.% \footnote{The vector particle is normally taken to be the $\omega$ meson, but for continuity with the earlier discussion, I will call it the $\rho$ meson.} Taking the short-distance couplings to be much stronger than the pion $(\alpha_\sigma,\alpha_\rho)=(7,14.65)$, the modified effective ranges $\widetilde{r}_n$ of the data can also be reproduced.~\cite{us2} This three-Yukawa model is reminiscent of the Bonn potential,~\cite{mach} which is known to model the data well. Applying the NN EFT to observables produced from Eq.~(\ref{3yuks}) indeed leads to a breakdown around $m_\sigma/2\sim300$~MeV. I can verify the breakdown is not set by the two-pion threshold $2m_\pi$ within this model by varying the pion and scalar masses. The $\sigma(600)$ particle is a common parameterization in nuclear physics for two-pion exchange contributions. The large scalar coupling indicates that including such effects through chiral symmetry might require the modified effective range expansion. Such an analysis could improve the EFT radius of convergence.% \footnote{A preliminary calculation, inspired by discussions at this workshop, seems to indicate this is true.~\cite{us3}} It might seem odd that two-pion effects could be strong, since they should be weak according to chiral symmetry.~\cite{lepage} However, there are other instances where two-pion effects have proven to be strong. The pion scalar form factor, defined by \begin{equation} \langle \pi^a(p) \pi^b(p') | {\hat m} \left( \bar{u} u + \bar{d} d \right) | 0 \rangle = \delta^{ab} \sigma_\pi \Gamma(s), \qquad\qquad s= (p+p')^2, \end{equation} has been shown to have poor convergence in chiral perturbation theory,~\cite{meiss} breaking down just above threshold at $\sqrt{s}\simeq 400$~MeV. This comes about from infrared effects that require the summation of $(\ln s)^n$ terms to all orders. Carrying out this summation does lead to improved results and the generation of a ``resonance''-like shape in the dispersion analysis just like a scalar resonance around $600$~MeV.~\cite{meiss} The same phenomenon could occur in NN scattering and requires further study. \begin{figure} \begin{center} \leavevmode \epsfxsize=3in \epsffile{3Yuk.eps} \end{center} \caption{\label{sig}The three-Yukawa model Eq.~(\protect\ref{3yuks}) with different values for the scalar mass, but $a$, $r_0$, and $m_\rho-m_\sigma$ fixed. This shows that the radius of convergence tracks with the changing $m_\sigma$.} \end{figure} \section{Red Herrings} The models developed to understand the breakdown scale better can also be used to shed light on other convergence issues. One feature of the three-Yukawa model is that it can reproduce the somewhat large effective range in NN scattering, which is $r_0\sim2/m_\pi$. Any reasonable choices of parameters for the two-Yukawa model can only give $r_0$ around $1/m_\pi$. Therefore, it is possible that the lower radius of convergence of the three-Yukawa model might have nothing to do with two-pion effects, but come from some type of correlation with the effective range $r_0$ instead. Choosing different parameters in Eq.~(\ref{3yuks}), I can keep $a$, $r_0$, and the mass difference $m_\rho-m_\sigma$ fixed while changing $m_\sigma$ from $500$~MeV to $1500$~MeV. The results are shown in Fig.~\ref{sig}, confirming that the radius of convergence is set by $m_\sigma/2$ alone. Recently, the convergence of $Q$-counting has been challenged within the context of a low-energy theorem.~\cite{cohen} The point can be made by defining the observable \begin{equation} S(p^2) \equiv p\cot\delta - \left( - \frac1{a} + \frac12 r_0 p^2 \right)\, , \end{equation} and taking $p=|1/a|$. Then $Q$-counting has two types of corrections: \begin{equation} \frac1{a \Lambda_{{\rm NN}}} \ll 1\, , \qquad\qquad\qquad \frac{m_\pi}{\Lambda_{{\rm NN}}} \sim \frac12\, . \label{exppar} \end{equation} It was shown that the $Q$-counting prediction for $S(p^2)$ at NLO is very far from the data. This can also be seen in a model. For an analytic understanding, I will use the two delta-shell potential of Section~\ref{fits}. Expanding in $Q$, \begin{eqnarray} S\left(\frac1{a^2}\right) &=& \left( -22.9+40.2-23.3+7.9+ \ldots\right) \times 10^{-7} m_\pi\, ,\\ &&\quad \mbox{\sc nlo} \quad\; \mbox{\sc nnlo} \quad \mbox{\ldots} \nonumber \end{eqnarray} one finds the NNLO term is larger than the NLO term. Only after a few more orders does the series start to converge. Taking the exact answer, I can also expand in $Q$ to see what is happening analytically. These expressions are messy and can be simplified by observing that even though $m_\pia ={\cal O}(1)$ in $Q$-counting, it is actually of order 10 numerically. Expanding $S(1/a^2)$ in $(m_\pi a)^{-1}$ then should simplify the expression immensely without blurring its meaning, leading to \begin{eqnarray} S\left(\frac1{a^2}\right) &=& \frac{m_\pi}{(m_\pi a)^4} \left[ - \frac{m_\pi}{3 \Lambda_{{\rm NN}}} + \frac{m_\pi^2}{\Lambda_{{\rm NN}}^2} \left( \frac{13}{15} + \frac{8 \Lambda_{{\rm NN}}}{9 m_\rho} \right) + \ldots \right] \\ &&{}+ \frac{m_\pi}{(m_\pi a)^5} \left[ \frac{4 m_\pi}{15\Lambda_{{\rm NN}}} -\frac{m_\pi^2}{\Lambda_{{\rm NN}}^2} \left( \frac{68}{45} + \frac{4\Lambda_{{\rm NN}}}{9m_\rho} \right) + \ldots \right] + \ldots\, . \end{eqnarray} Terms in the second bracket and higher are suppressed by $(m_\pi a)^{-1}$ and can be neglected. Focusing on the first bracket of terms, it is clear that the expansion is in $m_\pi/\Lambda_{{\rm NN}}$ as expected for $Q$-counting. However, the leading term (which happens to be NLO for this particular observable) has a numerical factor of $\frac13$, which suppresses it down to the order of the NNLO term. Therefore, the $Q$ expansion of $S(p^2)$ works, but is slowly converging due to some inconvenient numerical factors as well as the size of the expansion parameter, Eq.~(\ref{exppar}). \section{Summary} Nonperturbative effective field theories are subtle to implement. Techniques used successfully in chiral perturbation theory, such as naive dimensional regularization and low-momentum fitting, do not produce the expected results in nonperturbative calculations. After a careful study using solvable models, new methods have been developed to deal with these complications.~\cite{PDS,us2,kap} Pions enter the calculation differently depending on whether power counting is done in the QCD scale ($\Lambda$-counting) or the chiral scale ($Q$-counting). In both cases a proper determination of the low-energy constants from the short-distance physics requires more than just a low-momentum fit. Utilizing the modified effective range expansion to remove the effect of the long-distance pions from the process of fitting, the $\Lambda$-counting radius of convergence moves beyond the pion scale.~\cite{us2} A constrained global fit minimized with respect to the modified effective ranges is required to achieve the same effect in $Q$-counting.~\cite{kap} Applying the new fitting procedures to ${}^1\!S_0$ NN scattering data results in a breakdown of about $300$~MeV. Models can be used to confirm this breakdown is associated with the same physics for both power counting procedures. Using a three-Yukawa model, it was argued that including two-pion exchange will extend the EFT radius of convergence. It is not unreasonable to assume this effect is strong since a similar situation occurs in the pion scalar form factor.~\cite{meiss} Finally, the convergence of $Q$-counting was investigated using an exactly solvable model. It was shown that some observables could have inconvenient numerical factors, such as occur in the low-energy theorem,~\cite{cohen} which suppress the leading orders and mask the behavior of the expansion. However, analytical results show that this expansion does converge, albeit slowly. \section*{Acknowledgments} The work presented in this talk was done in collaboration with R.~J.~Furnstahl and D.~B.~Kaplan. I would like to thank the organizers, P.~Bedaque, M.~J.~Savage, R.~Seki, and U.~van Kolck, for an enjoyable and productive workshop and the Institute of Nuclear Theory at the University of Washington for its hospitality. This work was supported by the National Science Foundation under Grants No.\ PHY--9511923 and PHY--9258270. \section*{References}
\section{Introduction} The majority of the spin-down energy $\dot{E}$\ of a pulsar is carried away by a relativistic wind which is a combination of Poynting and kinetic-energy flux \cite{rg74,mic82}. The exact contributions to the wind from the magnetic and particulate components is uncertain and has only been determined for the Crab \cite{kc84} and PSR B1957+20 \cite{kpeh92}. In contrast, the characteristic coherent pulsed radio emission represents only a small fraction of the total energy liberated. Particles emerging from a pulsar magnetosphere have a null pitch angle and thus the relativistic wind is not directly observable. However, at a shock interface these pitch angles are randomised and produce synchrotron emission providing an excellent diagnostic of the wind (e.g. Frail et al. 1996). In the Crab Nebula the pulsar's wind is confined in the high pressure environment of the (invisible) surrounding supernova remnant (SNR) and a compact ``plerion'' is observed \cite{wp78}. Outside this high-pressure environment, the pulsar wind experiences the much lower pressure of the general interstellar medium (ISM). This is expected to result in the formation of very large ``ghost remnants'' \cite{bopr73}, an example of which is yet to be discovered \cite{ccgm83}. However, if the ram pressure resulting from the motion of the pulsar relative to the ISM is high, the interaction with the ambient medium may take the form of a bow-shock nebula. These nebulae are collectively known as pulsar wind nebulae (PWNe). Examples of plerions and bow-shock nebulae have been observed via their emission at optical (H$\alpha$) (e.g. Kulkarni \& Hester 1988\nocite{kh88}; Bell, Bailes \& Bessell 1993\nocite{bbb93}), radio (e.g. Frail et al. 1996\nocite{fggd96}), and X-ray (e.g. Wang, Li \& Begelman 1993\nocite{wlb93}) wavelengths. In general their radio emission is distinguished by having a flat spectral index and significant linear polarization. At present radio nebulae associated with pulsars are rare, with only seven examples known. Discovering more PWNe which emit at radio wavelengths is of great interest. The spectral, polarization and energy properties of such nebulae provide us with valuable information on the energy and particle composition of the wind, the local ambient ISM density and the nebular magnetic field. For a bow-shock nebula the cometary morphology and associated trail can constrain both the transverse velocity of the pulsar and the direction of that motion. This may be critical when considering possible SNR associations where the pulsar is outside the remnant. This paper reports on a search for radio PWNe around five southern pulsars. It was partially inspired by the present sparsity of radio PWNe and also by the recent VLA survey for PWNe carried out by Frail \& Scharringhausen (1997, hereafter FS97)\nocite{fs97}. Their source list was made up of 35 candidates chosen for their high space velocity and/or high $\dot{E}$. To avoid contamination of a possible PWN by the pulsar itself they decided to observe at 3~cm where the flatter spectrum of any PWN should cause it to dominate its associated pulsar. Unfortunately, however, FS97 detected no new nebulae in their search. As those radio PWN already observed have $\dot{E}$\ $> 10^{35}$\,erg\,s$^{-1}$, while the majority of the pulsars searched had $\dot{E}$\ $< 10^{35}$\,erg\,s$^{-1}$, FS97 concluded that only young, energetic pulsars produce radio-bright PWNe. However, their choice of observing frequency and array configuration meant they could not have detected PWNe larger than 20 arcsec, and also had poor sensitivity to low surface brightness emission on smaller scales. They also could not tell if unresolved emission corresponded to compact PWNe or just to the pulsar itself. In this experiment, we have used pulsar gating on the Australia Telescope Compact Array (ATCA) to remove the pulsed emission. This allows us to detect compact PWNe ``hidden'' behind the pulsar itself. It also enables us to observe at lower frequencies and resolution, resulting in a significantly more sensitive search. Our 20\,cm observations have a typical (1\,$\sigma$) sensitivity of 70\,$\mu$Jy\,beam$^{-1}$ at $\sim$12 arcsec resolution, compared to FS97's 40\,$\mu$Jy\,beam$^{-1}$ at $0.8$ arcsec. Hence, for a typical PWN spectral index $\alpha = -0.3$ ($S_{\nu} \propto \nu^{\alpha}$), we are able to detect a PWN of surface brightness \mbox{200 times fainter} than could FS97. Our observations are also sensitive to PWNe on scales up to $\sim$3 arcmin, $\sim$80 times the size of the largest sources which FS97 could detect. These observations were supplemented by Molonglo Observatory Synthesis Telescope (MOST) observations at 36\,cm allowing us to probe spatial scales up to 80 arcmin. This search targetted five southern pulsars (PSR B0906$-$49, PSR B1046$-$58, PSR B1055$-$52, PSR J1105$-$6107, PSR B1610$-$50). A nebula associated with PSR B0906$-$49\ was successfully detected and is discussed in detail by Gaensler et al. (1998)\nocite{gsfj98b}. We here report non-detections of PWNe towards the remaining candidates. Candidate selection is discussed in Section 2, while in Section 3 we discuss the observations and the pulsar-gating mode which we employed. In Sections 4 and 5 we present our results and go on to consider the corresponding implications. \section{The Sample} Because the ATCA is an East-West synthesis telescope, an image of each pulsar requires a full 12-hour synthesis, and on a reasonable timescale it is not possible to undertake a survey on the scale of that made by FS97. We therefore chose our sample of 5 pulsars based on a combination of their characteristic age ($\tau_{c} = P/2\dot{P}$), $\dot{E}$, and reported detections of a PWN at other wavelengths. The characteristics of the four pulsars, including their Galactic longitude, $l$, and latitude, $b$, rotational period, characteristic age, $\dot{E}$, dispersion measure, and the distance implied by the dispersion measure \cite{tc93}, are given in Table \ref{char}. We now consider each pulsar in detail. \begin{table*} \begin{center} \caption{ATCA observations of PWN candidates.} \begin{tabular}{rrrrrrrr} \hline \hline \multicolumn{1}{c}{Source}&\multicolumn{1}{c}{Date}&\multicolumn{1}{c}{Secondary}& \multicolumn{1}{c}{B$_{\rm max}$}&\multicolumn{2}{c}{$\nu$\,(GHz)}&\multicolumn{1}{c}{$t_{res}$$^{\rm a}$} \\ \multicolumn{1}{c}{}&\multicolumn{1}{c}{}&\multicolumn{1}{c}{Calibrator}&\multicolumn{1}{c}{(m)}&\multicolumn{1}{c}{1}&\multicolumn{1}{c}{2}&\multicolumn{1}{c}{(ms)}\\ \hline B1046--58 & 1997 May 13 & PKS B1215-457 & 5969 & 1.344 & 2.240 & 3.9\\ B1055--52 & 1997 May 14 & PKS B1215-457 & 5969 & 1.344 & 2.240 & 6.2\\ J1105--6107& 1995 June 1 & PKS B1036-697 & 750 & 1.376 & $\ldots$ & 7.9\\ J1105--6107& 1995 May 16 & PKS B1036-697 & 1485 & 1.376 & $\ldots$ & 7.9\\ B1610--50 & 1997 May 10 & MRC B1613-586 & 5969 & 1.344 & 2.496 & 8.0(7.2)\\ \hline \end{tabular} \label{obs} \end{center} (a) The values in parentheses corresponds to the effective time resolution at 13\,cm for PSR B1610$-$50. For all other pulsars it does not differ from the effective time resolution at 20\,cm. \end{table*} \begin{figure} \leavevmode\epsfig{file=onoff.ps, angle=-90, width=\linewidth} \caption{Two {\tt CLEAN}ed images of the field around PSR B1046$-$58\ are shown. Above all bins have been used; below only the off pulse bins are included. The ease of identification of any non-pulsed emission are immediately apparent. No PWN is discernible in the off-pulse image on any spatial scale, a conclusion which can only be drawn when pulsar gating is applied.} \label{onoff} \end{figure} \subsection{PSR B1046$-$58} PSR B1046$-$58\ was discovered in a high-frequency survey of the Galactic plane \cite{jlm+92}. It is young, $\tau_c$\ $\approx$ 20\,kyr, and has the 11th highest $\dot{E}$\ of all known pulsars. Despite its youthfulness it is not near any known SNR. Its high $\dot{E}$\ and relative proximity (2.5$\pm$0.5\,kpc, Johnston et al. 1996\nocite{jkww96}) mean that it ranks in the top 10 pulsars when considering the spin-down energy flux at the Earth. Thus, it is an excellent candidate for having detectable high-energy emission (e.g. Fierro 1995\nocite{fie95}) and is indeed coincident with the {\em EGRET} source 2EG J1049-5847 \cite{nab+96} and pulsations have probably been detected \cite{klp+98}. There is also X-ray emission in the direction of the pulsar which has been detected by both {\em ASCA} \cite{kt96,kts98a} and {\em ROSAT} \cite{bt97}. The {\em ASCA} results suggest that this emission is extended and may be from a pulsar-powered nebula. \subsection{PSR B1055$-$52} PSR B1055$-$52\ is not particularly young, energetic nor rapidly rotating. However it is one of only a small number of pulsars which are detected across the whole electromagnetic spectrum and one of only seven pulsars known to pulse in high-energy $\gamma$-ray s. It is a bright X-ray source \cite{ch83} with the soft photons modulated at the pulse period \cite{of93} and is the pulsar which converts the largest fraction of its spin-down energy into pulsed $\gamma$-ray\ emission \cite{tbb+99}. Recent {\em HST} observations by Mignani, Caraveo \& Bignami (1997)\nocite{mcb97} were successful in detecting the optical counterpart. The optical emission is consistent with thermal emission from the neutron star surface. The {\em Einstein} detection of PSR B1055$-$52\ showed an apparently resolved source, suggesting that the X-ray emission was from a PWN \cite{ch83}. However {\em ROSAT} observations made by \"Ogelman \& Finley (1993) indicate no extended structure. Subsequent {\em ASCA} observations in the 2--8\,keV X-ray band show a clumpy 20 arcmin ring around the pulsar which has also been interpreted as a PWN \cite{ssg+97}. Although not conclusive these X-ray results suggest PSR B1055$-$52\ may be embedded in a PWN. Combi, Romero \& Azc\'arate (1997)\nocite{cra97} have searched for an associated radio PWN using a 30\,m dish at 20~cm. They detected a 66\hbox{$^\prime$}$\times$126\hbox{$^\prime$}\ radio source overlapping the pulsar's position and the proposed X-ray nebula. Comparing this emission with a similar structure they see in the 408\,MHz (74\,cm) all-sky survey of Haslam et al. (1982)\nocite{hks+82} they measured a non-thermal spectrum and suggested that the source may represent synchrotron emission from a PWN powered by PSR B1055$-$52. \subsection{PSR J1105$-$6107} PSR J1105$-$6107\ is both young and energetic \cite{kbm+97}. Like PSR B1046$-$58\ its high $\dot{E}$\ suggests that it may be a $\gamma$-ray\ source and it is coincident with 2EG J1103--6106. However Kaspi et al. (1998)\nocite{klp+98} find no evidence for $\gamma$--ray pulsations attributable to PSR J1105$-$6107. This pulsar is also near the SNR G290.1--00.8 and an association has been proposed by Kaspi et al. (1997)\nocite{kbm+97}. For the age and distance given in Table~\ref{char}, the association with G290.1--00.8 implies a projected velocity for the pulsar of $v_{\rm t} \approx 650$\,km\,s$^{-1}$ in order to have reached its present location well outside the SNR. X-ray observations with {\em ASCA} indicate the presence of an unpulsed source at the pulsar's position which is interpreted as a PWN \cite{gk98b}. \subsection{PSR B1610$-$50} PSR B1610$-$50\ is the third youngest known pulsar in the Galaxy. Its youth suggests that it be associated with a SNR and indeed an association with the nearby SNR G332.4+00.1 (Kes~32) has been claimed by Caraveo (1993)\nocite{car93}. The pulsar is $\sim$12 arcmin from the center of the SNR, so an association therefore requires a transverse velocity for the pulsar, $v_{\rm t} \approx 460\times d_{\rm kpc}$\,km\,s$^{-1}$, where $d_{\rm kpc}$ is the distance to the pulsar in kpc. Its distance is quite uncertain, but taking the minimum remnant/pulsar distance, $d_{\rm kpc} \sim 5$ (e.g. Johnston et al. 1995\nocite{jml+95}) gives a distinctly high implied velocity, $v_{\rm t} = 2300$\,km\,s$^{-1}$. Such a large velocity should result in a large ram pressure with the ISM, and thus a radio-bright PWN with a cometary tail indicating the pulsar's motion might be expected \cite{fk91}. As is typical for a young pulsar, PSR B1610$-$50\ shows excessive timing noise and glitches, and thus its position from pulse timing analysis is not well known \cite{jml+95}. However an added advantage of the gated observations used here is that the pulsar's position can be accurately determined. \section{Observations and Data Reduction} Observations made with the ATCA \cite{fbw92} are described in detail in Table \ref{obs} where B$_{\rm max}$ is the maximum baseline and $\nu$ is the central observing frequency. Observations were made simultaneously at approximately 20~cm and 13~cm, except for PSR J1105$-$6107\ for which only 20~cm data were obtained. A total bandwidth of 128\,MHz, divided into 32 channels, and all Stokes parameters was recorded at both wavelengths. All observations used the ATCA's pulsar gating mode, which divides the visibilities into a maximum of 32 time bins per pulse period, with a minimum width of $\sim$2.5\,ms. A polynomial description of the apparent pulse period is then used to fold these bins online. Summed visibilities were then recorded every 40\,s. The more rapid rotation rate of PSR J1105$-$6107\ meant that only 8 bins were used, while all 32 bins were recorded for the remaining pulsars. The effective time resolution, $t_{res}$ for both frequencies is given in Table \ref{obs}. The dispersion measure smearing inside each 4\,MHz channel was less than or approximately equal to the sampling time in all cases except for PSR B1610$-$50\ at 20~cm (both 20\,cm and 13\,cm resolutions for PSR B1610$-$50\ are therefore given in Table \ref{obs}). For PSR B1046$-$58, PSR B1055$-$52\ and PSR B1610$-$50, the pulsar was offset from the phase center by a few arcminutes. Observations of PSR J1105$-$6107\ were centered on the SNR G290.1--01.8, so that the pulsar was approximately 22 arcmin away from the field center. For all sources the flux density scale was determined by observations of PKS B1934--638 using the revised flux density scale of Reynolds (1994)\nocite{rey94}. The antenna gains were determined by observations, every 40 minutes, of the secondary calibrators given in Table \ref{obs}. Data reduction, including editing and calibrating, was carried out using the {\tt MIRIAD} package and standard techniques \cite{sk98}. Once calibrated, the data were dedispersed to the center frequency of the observations using the dispersion measures (DMs) given in Table \ref{char}. An image of each field was made, deconvolved using the {\tt CLEAN} algorithm \cite{cla80}, and restored using a Gaussian beam with dimensions appropriate for the diffraction limit (see $\theta_{min}$ in Table \ref{limits}) and finally a correction for primary beam attenuation was made. Each pulsar was identified by subtracting, after dedispersion and appropriate scaling, an image made from the first bin alone from an image made using all bins. As the pulse was the only unique feature in any bin it is the only source in the image\footnote{Noting that if the strongest section of the pulse is in the first bin the pulsar will appear negative}. The pulsar's position was determined by fitting a point source to the on-pulse $u-v$ data, and the pulse profile was extracted from the entire data-set at that position. The off-pulse bins were used to re-image the field and generate an image which contained no pulsed emission. Any underlying PWN emission can then be identified. Fig. \ref{onoff} compares the effect of using all bins when observing PSR B1046$-$58\ (i.e. equivalent to observing without gating) with that of using only the off-pulse bins. Our ATCA observations are only sensitive to a maximum spatial scale of $\sim$3 arcmin. To search for more extended structures we used data from the MOST \cite{rob91}, an East-West synthesis telescope operating at a fixed wavelength of 36~cm. The MOST images all spatial scales between 43 arcsec and $\sim$80 arcmin with a typical sensitivity (for a 12\,h synthesis) of 1\,mJy\,beam$^{-1}$, and is therefore particularly sensitive to large-scale structure. These observations did not use pulsar gating. Data from the MOST Galactic plane survey \cite{gcly99} were used for PSR B1046$-$58, PSR J1105$-$6107\ and PSR B1610$-$50, while a separate observation was made of PSR B1055$-$52\ on 1998 Apr 4. \begin{figure} \leavevmode\epsfig{file=profiles.ps, width=10cm,angle=-90} \caption{The 20\,cm pulse profiles for the pulsars, as labelled, detected in these observations. Profiles at 13\,cm looked similar. Thirty two time bins were used and the time resolution is given in Table \ref{obs}. In all three cases there are sufficient off-pulse bins to generate a high signal-to-noise off-pulse map. Absolute timing was not carried out and the reported pulse phase is therefore arbitrary.} \label{profiles} \end{figure} \section{Results: Pulsed Emission} Pulsations from PSR B1046$-$58, PSR B1055$-$52\ and PSR B1610$-$50\ were successfully detected in our observations and their 20~cm pulse profiles are presented in Fig. \ref{profiles}. These profiles, within the limited resolution, are consistent with the known pulse shapes and similar profiles were obtained at 13\,cm. In all cases the pulsed emission can be clearly separated from any off-pulse contribution. Flux densities and the interferometric positions for all three detected pulsars are given in Table \ref{fluxes}. \begin{table*} \begin{center} \caption{Observed parameters of the detected pulsars.$^{\rm a}$} \begin{tabular}{rrrrrr} \hline \hline \multicolumn{1}{c}{Name} & \multicolumn{2}{c}{Interferometric Position} & \multicolumn{3}{c}{Flux Density (mJy)} \\ &\multicolumn{1}{c}{R.A. (2000)} & \multicolumn{1}{c}{Dec. (2000)} & 36\,cm & 20~cm & 13~cm \\ \hline B1046--58 & 10:48:12.6(3) & -58:32:03.75(1) & 10(3) & 9.4(5) & 7.0(1) \\ B1055--52 & 10:57:58.9(9) & -52:26:56.1(2) & 15(2) & 8.7(2) & 3.5(1) \\ B1610--50 & 16:14:11.6(8) & -50:48:01.90(9) & $\dots$ & 4.7(2) & 1.0(1) \\ \hline \end{tabular} \label{fluxes} \end{center} (a) The values in parentheses correspond to the 1\,$\sigma$ errors in the final digit.\end{table*} The flux of PSR B1046$-$58\ at 20~cm is consistent with that in the pulsar catalogue \cite{tmlc95}, while PSR B1610$-$50\ has a 20~cm flux approximately twice that given in the catalogue; no catalogue value for the 20~cm flux is given for PSR B1055$-$52. A confusing thermal source, discussed below, is located near the position of PSR J1105$-$6107\ and prevented its detection. The sensitivity limit for detection was very close to the known 20~cm flux density of PSR J1105$-$6107, 1.8\,mJy \cite{kbm+97}. The position (Table \ref{fluxes}) of PSR B1055$-$52\ matches the timing position to within the errors and our position for PSR B1046$-$58\ is only $\approx 2.3$ arcsec different in declination. The position of PSR B1610$-$50\ is ill-determined from timing and our improved position, RA (J2000) 16$^{\rm h}$14$^{\rm m}$11\ras6, Dec (J2000) -50$^{\circ}$48\arcmin01\arcsecdot90, is approximately half an arcminute away from the nominal timing position of Johnston et al. (1995)\nocite{jml+95}. \section{Results: Off-pulse Emission} No off-pulse emission, either extended or unresolved, was detected from any of the four pulsars discussed here. The 1\,$\sigma$ surface brightness limits, $\sigma_{rms}$, for these non-detections are given in Table \ref{limits}. These limits are determined by measuring the rms variation in the off-pulse image in the vicinity of the pulsar. Also shown are the minimum, $\theta_{min}$, and maximum, $\theta_{max}$, angular scales to which we were sensitive at 20~cm. Two observations were made of PSR J1105$-$6107\ one of which used a more compact configuration thereby increasing the angular size to which we were sensitive. \begin{table*} \begin{center} \caption{Limits on nebular and unpulsed flux densities.} \begin{tabular}{rrrrrrrrrr} \hline \hline \multicolumn{1}{c}{Source} & \multicolumn{3}{c}{$\sigma$\,(mJy/beam)} & \multicolumn{1}{c}{$\theta_{\rm min}$} & \multicolumn{1}{c}{$\theta_{\rm max}$} & \multicolumn{1}{c}{L$_{\rm R}$} & \multicolumn{1}{c}{L$_{\rm R}$/$\dot{\rm E}$} & \multicolumn{2}{c}{$\sigma_{\rm off}$/S}\\ & 36~cm & 20~cm & 13~cm & (\hbox{$^{\prime\prime}$}) & (\hbox{$^\prime$}) & (erg s$^{-1}$) & ($10^{-6}$)& 20\,cm & 13\,cm \\ \hline B1046--58 & 0.8 & 0.14 & 0.08 & 10 & 3.2 & 3$\times10^{29}$ & 0.15& 0.015 & 0.011\\ B1055--52 & 1.2 & 0.08 & 0.16 & 10 & 3.2 & 4.5$\times10^{28}$ & 1.5 & 0.009 & 0.05 \\ J1105--6107& 12 & 0.3 & $\ldots$ & 20 & 15.6& 2.5$\times10^{30}$ & 1.6 & $\ldots$ & $\ldots$ \\ B1610--50 & 5.0 & 0.2 & 0.16 & 12 & 3.2 & 3.5$\times10^{30}$ & 1.4 & 0.04 & 0.15 \\ \hline \end{tabular} \label{limits} \end{center} \end{table*} The lack of any unpulsed emission in the vicinity of PSR B1046$-$58\ is starkly apparent in Fig. \ref{onoff}. The 5\,$\sigma$ limits for any undetected point-like nebula are 0.7\,mJy and 0.4\,mJy at 20 and 13~cm respectively. Fig. \ref{xrayneb} shows a 20\,cm off-pulse image of the field surrounding PSR B1055$-$52\ with the {\em ASCA} X-ray contours in the energy band 0.5--2.0 keV overlayed. The position of the pulsar is marked by the cross and the contours corresponding to the X-ray emission from the pulsar and the three ``clumps'' \cite{ssg+97} are clearly seen. There is no indication of any extended or point-like radio emission detected at the position of the pulsar. Radio emission is seen at the position of at least two of the three clumps, however it is clearly resolved into a number of discrete point sources and does not appear to form part of a large extended structure. The location of PSR J1105$-$6107\ relative to SNR~G290.1--0.8 is indicated by the cross in the MOST 36\,cm image in the left hand panel of Fig. \ref{vkfig}. Similarly the pulsar is marked in the higher resolution 20\,cm ATCA image of the immediate region around the pulsar's location. Neither of these images have been gated and the pulsar is undetected at both frequencies. An extended source, whose center is some 4 arcmin to the North of the pulsar, is clearly seen in the 20\,cm image and is also detected at 36\,cm. As this emission is diffuse and is coincident with an IRAS 60\,$\mu$m source it is probably thermal and not associated with the pulsar. The pulsar is un-detected but the 5\,$\sigma$ limit on any possible PWN radio emission was only 1.5\,mJy at 20~cm. \begin{figure} \leavevmode\epsfig{file=1055olay.ps,angle=-90, width=\linewidth} \caption{Radio and X-ray observations of PSR B1055$-$52. The grey-scale corresponds to a 20\,cm off-pulse image, overlayed with X-ray contours from the {\em ASCA} 0.5--2\,keV energy band. The X-ray contours are at 50\%, 60\%, 70\%, 80\% and 90\% of the peak value. The position of the pulsar is labelled by the cross and the three clumps of Shibata et al. (1997) are indicated. The X-rays associated with the pulsar are clearly seen. There is no evidence for unpulsed radio emission at or around the position of the pulsar.} \label{xrayneb} \end{figure} The 36\,cm MOST image of the complicated field surrounding PSR B1610$-$50\ is shown in Fig. \ref{full}. SNR~Kes 32 is seen to the north-east while SNR~G332.0+00.2 is to the south-west. Using our ATCA data we also show a higher resolution image of the region immediately surrounding the pulsar. After gating out the pulsar's emission we detect no unpulsed emission except for a filled-center clump located some 2 arcmin north of the pulsar which we designate G332.23+00.2. No linear polarization was detected from G332.23+00.2 and we measure flux densities of 0.20$\pm0.05$\,Jy (36~cm) and 0.20$\pm0.02$\,Jy (20~cm), suggesting a flat spectral index, while it was largely resolved out in our 13~cm observations. \section{Discussion} We have failed to detect any radio-bright PWNe emission over a wide range of angular scales and down to good sensitivities around PSRs B1046$-$58, B1055$-$52, J1105$-$6107\ and B1610$-$50. These non-detections have interesting implications for all of these pulsars and we will consider them in detail below. Furthermore the gating method employed here also allows us to put a strong limit on the off-pulse emission from the pulsar itself. In Table \ref{limits} we give the limits on the ratio of the off-pulse rms noise to the on-pulse flux density. For PSR B1046$-$58\ and PSR B1055$-$52\ our limits are strong, and indicate that for these young pulsars any unpulsed radio emission from the neutron star itself is less than 2\% of the pulsed emission (c.f. Hankins et al. 1993)\nocite{hmnp93}. For PSR J1105$-$6107\ and PSR B1610$-$50, we might have expected compact, ram-pressure confined PWNe, resulting from the high velocities inferred through associations with nearby SNRs. The emission from such nebulae may be located quite close to the termination shock radius, $r_{\rm s}$. For example we can derive an expression for the expected angular size of this shock (e.g. Cordes 1996)\nocite{cor96} near PSR B1610$-$50\ for a spherical wind; \begin{equation} \theta = \frac{0\arcsecdot15}{d_7v_{1000}}\sqrt{\frac{f}{n}}, \end{equation} where $v = v_{1000}\times1000$\,km\,s$^{-1}$ is the pulsar's velocity, $d = d_7\times7$\,kpc is the distance, $n$ is the density of the local ISM in cm$^{-3}$ and $f$ is the efficiency of conversion of spin-down energy to wind energy. Hence emission near $r_{\rm s}$ would be unresolved in our observations under the assumption of high-velocity. \begin{figure*} \leavevmode\epsfig{file=1105_ma.ps, angle=-90, width=\linewidth} \caption{Radio observations of PSR J1105$-$6107. Left: A 36~cm MOST image of the field, including the SNR~G290.1--0.8. Right: A higher resolution, 20~cm, ATCA image of the immediate region surrounding the pulsar. The ATCA synthesised beam is shown at lower right. No pulsar gating has been applied to either image, and the pulsar position is indicated by the cross. The thermal emission directly to the North of the pulsar can clearly be seen in the ATCA image.} \label{vkfig} \end{figure*} In fact, if we assume that all the pulsars in our sample are moving with a typical pulsar velocity, 250\,km\,s$^{-1}$ (e.g. Hansen \& Phinney 1997\nocite{hp97}) and in ambient medium of density $n\approx0.1$\,cm$^{-3}$, PWN around them will be unresolved in our 20~cm data. Similarly to FS97, using our non-detections we can then determine an upper limit on the radio luminosity, L$_{\rm R}$, for these compact nebulae. Assuming a Crab-like spectral index (noting that the spectral index of the PWN associated with PSR B0906$-$49\ is steeper); \begin{equation} L_{\rm R} = 4.74\times10^{28}d^2_{kpc}S_{20}\,{\rm erg\,s}^{-1}, \label{rlum} \end{equation} where $d_{\rm kpc}$ is the distance to the nebulae and $S_{20}$ is the 20\,cm flux limit. The L$_{\rm R}$ for each non-detection, based on the 5\,$\sigma$ limiting flux density at 20\,cm, is given in Table \ref{limits}. Comparing these maximum L$_{\rm R}$ values with the known $\dot{E}$\ we find that the pulsars in our sample have efficiencies (see Table \ref{limits}) which lie in or below the peak of the distribution shown by FS97 at L$_{\rm R}$/$\dot{E}$$\sim10^{-6}$. Thus for an unresolved nebula the efficiencies are well below those at which radio PWNe appear to be detected. An alternative to this implied low efficiency for the lack of PWN could simply be the fact that the PWN are resolved but are too faint to see. Both PSR B1046$-$58\ and PSR B1610$-$50\ have similar ages, spin-down energies and surface magnetic fields to PSR B1757--24 \cite{fggd96} and PSR B1853+01 \cite{fk91} which are known to have associated PWN. If we therefore assume that these pulsars all have similar wind characteristics and that conditions at the shock with the ISM are the same, then the ratio of the resultant PWN's radio luminosity to the spin-down energy should also be similar. Using Equation \ref{rlum}, together with the distances and spin-down energies given in Table \ref{char} and L$_{\rm R}$/$\dot{E}$$\sim2\times10^{-4}$ (e.g. PSR B1757--24; FS97), we find that PWN associated with PSR B1046$-$58\ and PSR B1610$-$50\ should have S$_{20} = 1$ and 0.1\,Jy respectively. To reconcile these expected flux densitys with the observed surface brightness limits these nebulae must be larger than a certain radius. Assuming circular nebulae with uniform surface brightness our 3\,$\sigma$ flux-density limits correspond to PWNe of minimum radii 8 and 3\,pc respectively\footnote{The MOST offered the best sensitivity limit for PSR B1046$-$58. Hence, assuming a Crab-like spectral index, the 20\,cm flux limit was converted to a 36\,cm flux limit for comparison with the MOST upper limit}. Now consider that this minimum radius corresponds to the radius of a static PWN \cite{aro93}; \begin{equation} R_s = \left(\frac{\dot{E}}{4\pi\rho_0}\right)^{1/5}t^{3/5}{\rm cm}, \label{rstatic} \end{equation} where $\rho_0$ is the density of the surrounding medium in g\,cm$^{-1}$\,s$^{-2}$ and $t$ is the age, we can derive a number density of the surrounding medium, $n_0$. Around PSR B1046$-$58\ $n_0\lta2.2\times10^{-3}$\,cm$^{-3}$ and for PSR B1610$-$50\ $n_0\lta1.0\times10^{-2}$\,cm$^{-3}$, where we have assumed a purely hydrogen ISM. If instead we consider the case where the winds are in pressure balance at the aforementioned radii due to ram pressure then we find that $n_0v^2\lta0.4$\,cm$^{-3}$\,km$^{2}$\,s$^{-2}$ for PSR B1046$-$58\ and $n_0v^2\lta2$\,cm$^{-3}$\,km$^{2}$\,s$^{-2}$ for PSR B1610$-$50. Thus for reasonable pulsar velocities the implied density of the ISM is exceptionally low and we conclude that these nebulae are probably not ram pressure confined. In this case the pulsar velocity, $v$, cannot be significantly greater than the ISM shock velocity and thus, \begin{equation} v \approx \dot{R}_S = \left(\frac{3}{5}\right)\left(\frac{\dot{E}}{\rho_0t^2}\right)^{1/5}\,{\rm cm\,s}^{-1}. \label{time} \end{equation} Equating $t$ to the characteristic age of the pulsar and using our limits on $n_0$ above, we obtain significant constraints on the maximum space velocity of PSR B1046$-$58\ and PSR B1610$-$50\ of 480\,km\,s$^{-1}$ and 450\,km\,s$^{-1}$, respectively. \subsection{PSR B1046$-$58} \begin{figure*} \leavevmode\epsfig{file=1610_ma.ps, width=\linewidth} \caption{Radio observations of PSR B1610$-$50. Left: A 36~cm MOST image of the field, including the SNRs Kes 32 and G332.0+00.2. Right: A higher resolution, 20~cm, ATCA image of the immediate region surrounding the pulsar. This image was formed by combining our ATCA data with other, archival, observations of the region to give improved $u-v$ coverage and sensitivity. No pulsar gating has been applied to this image, and the pulsar is indicated. A knot of emission, G332.23+00.20, can be seen to the north. The ATCA synthesised beam is shown at lower right.} \label{full} \end{figure*} PSR B1046$-$58\ is a young pulsar which we have shown is probably in a low density region which is consistent with it having no associated SNR. Located near the pulsar there is a 100\,pc expanding gas shell centered on a pair of OB associations \cite{chty81}. Johnston et al. (1996) speculate that this shell may be related to the SN which formed PSR B1046$-$58. It is therefore probable that despite being very energetic, PSR B1046$-$58\ dwells in an environment which precludes the formation of a radio bright PWN. \subsection{PSR B1055$-$52} To properly compare our observations with the proposed extended nebula detected by Combi et al. (1997) we smoothed our data to $\sim$24 arcmin, the same resolution as their measurements. When smoothed, the distribution of point sources shown in Fig.~\ref{xrayneb} almost exactly reproduces the size and shape of the Combi et al. (1997) emission at 20~cm. Furthermore, the MOST data are sensitive to emission on scales as large as 80 arcmin, but show no evidence for extended radio emission. Thus we conclude that the source claimed as a PWN by Combi et al. (1997) is spurious, and is simply the effect of confusing sources observed with a small dish. The proposed X-ray PWN associated with PSR B1055$-$52\ is made of three clumps as indicated in Fig. \ref{xrayneb}. At least two of these three clumps coincide with clusterings of radio point sources seen in our 20~cm observations, and also with X-ray point sources seen in archival {\em ROSAT}\ data \cite{bec98}. We thus think it more likely that the {\em ASCA}\ clumps correspond to emission from unrelated background sources, rather than to a PWN associated with PSR B1055$-$52. To summarise, we argue that there is no evidence for either a radio or an extended X-ray PWN around PSR B1055$-$52. It may be that PSR B1055$-$52\ is converting the majority of its spin-down luminosity into high energy emission, rather than into a relativistic wind. PSR B1055$-$52\ may emit as much as 21\% (assuming emission into 1 steradian) of its spin-down energy at energies greater than 1\,eV \cite{tbb+99} which is significantly more than for the other pulsars in our sample. \subsection{PSR J1105$-$6107} Extensive thermal emission in the vicinity of PSR J1105$-$6107\ restricts the sensitivity limits which we can place on any potential radio PWN. If this thermal source is an H\,{\sc ii} region and lies along the line of sight to the pulsar then this may mean that the pulsar distance derived from its dispersion measure is significantly over-estimated. Extrapolating the power-law spectrum of the X-ray PWN ($\alpha =-0.8\pm0.4$; Gotthelf \& Kaspi 1998\nocite{gk98b}) through to radio wavelengths, we find that we expect a 20~cm flux density for any radio PWN of $\sim$0.5~Jy. This corresponds to a 5\,$\sigma$ detection of a $\sim$3 arcmin sized nebula in our 20~cm data. The expected radio flux falls below our detection threshold provided there is a break to a Crab-like spectrum ($\alpha=-0.3$) at a frequency above 10$^{14}$\,Hz. \subsection{PSR B1610$-$50} As shown in Fig.~\ref{full}, the clump of emission G332.23+00.2 is quite close to the pulsar. While its filled-centre morphology and flat spectrum are what might be expected from a PWN, the pulsar is well outside this source and there is no morphological evidence to suggest the two are associated. Rather, the non-detection of linear polarization from G332.23+00.2 suggests that it is an unrelated thermal region. The presence of strong IRAS 60\,$\mu$m emission nearby prevents a clear determination of whether G332.23+00.2 also emits at this wavelength. The lack of a PWN associated with PSR B1610$-$50\ indicates that the pulsar is located in a low-density environment and is moving with moderate velocity. The probability of a chance association along the line-of-sight to this pulsar is high, as this complex region contains several SNRs and pulsars (e.g. Johnston et al. 1995\nocite{jml+95}; Gaensler \& Johnston 1995\nocite{gj95b}). Furthermore, unlike the case of PSR~B1757--24 and G5.4-1.2 \cite{fk91}, there is no evidence of disruption of the SNR where the pulsar would have passed through the shell. From consideration of all these points, we strongly argue that the association by Caraveo (1993)\nocite{car93} of PSR B1610$-$50\ with Kes~32 is inconsistent with our data. Thus, even though PSR B1610$-$50\ is a very young pulsar, there are no obvious SNRs with which it might be associated. We do not consider this a problem, however, we believe it to be located in a medium with density $n_0\lta1.0\times10^{-2}$\,cm$^{-3}$ and since a large fraction of the ISM is believed to consist of low density coronal gas (e.g. McKee \& Ostriker 1977\nocite{mo77}), SNRs expanding into such regions will be undetectable (e.g. Kafatos et al. 1980\nocite{ksbg80}). \section{Conclusions} Using the excellent sensitivity to unresolved nebulae afforded by the gating mode of the ATCA and the greater spatial sensitivity of the MOST we have searched for PWNe around five pulsars. We successfully detected a new nebula associated with PSR B0906$-$49\ and failed to detect any extended or unresolved nebulae associated with the remaining candidates. We also derived strong limits on the unpulsed emission from PSRs B1046$-$58, B1055$-$52\ and B1610$-$50. These data argue strongly that the proposed ring-shaped X-ray nebula and associated radio nebula around PSR B1055$-$52\ are in fact due to emission from background sources. A lack of a nebula associated with PSR B1055$-$52\ combined with its observed high energy characteristics indicates that it probably converts most of its spin-down energy into high energy emission and not a relativistic wind. The lack of a nebular associated with PSR B1610$-$50\ clearly shows that it is not moving sufficiently fast to be associated with Kes\,32. PSR J1105$-$6107\ is located near the edge of a thermal region which will make identification of any radio PWN difficult. If the undetected nebulae around all our candidate pulsars are compact, then these pulsars convert less than 10$^{-6}$ of their spin-down energy into radio emission. Alternatively, assuming that PSR B1046$-$58\ and PSR B1610$-$50\ have similar winds to PSR B1757--24 and PSR B1853+01 we derive a maximum ISM number density for PSR B1046$-$58\ and PSR B1610$-$50\ of $n_0\lta2.2\times10^{-3}$\,cm$^{-3}$ and $n_0\lta1\times10^{-2}$\,cm$^{-3}$, respectively. The low-density is supported by the lack of any potential parental SNRs associated with these young pulsars. We also constrain their spatial velocities to be less than $\sim$450\,km\,$s^{-1}$. Our failure to detect nebulae around the young, and very energetic pulsars PSR B1046$-$58\ and PSR B1610$-$50\ indicates that a high spin-down rate is insufficient to generate a PWN. When combined with the very faint nebulae we did detect associated with the somewhat less energetic PSR B0906--49 these results highlight the role that the ISM plays in the production of such nebulae. The sensitivity limits achieved here would be unattainable without the use of pulsar gating and we now plan to employ this facility on the VLA to search a larger sample of pulsars. Such a sample will not only provide vital data on the wind parameters of a varied range of pulsars but will also supply valuable information on the ISM. \section*{Acknowledgments} We would like to thank Dale Frail for valuable discussions during the preparation of this work, Shinpei Shibata for providing the ASCA image of PSR B1055$-$52\ and Dick Manchester for supplying the pulsar ephemerides. The Australia Telescope is funded by the Commonwealth of Australia for operation as a National Facility managed by CSIRO. The MOST is operated by the University of Sydney with support from the Australian Research Council and the Science Foundation for Physics within the University of Sydney. BMG acknowledges the support of an Australian Postgraduate Award, and of NASA through Hubble Fellowship grant HF-01107-01-98A 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.
\section{Introduction} Exploding massive stars, and supernovae in particular, are known to be major sites for the production of a large variety of elements heavier than carbon. One of the few available ways to study the physics in the deep interior of such stars is the determination of the abundances of stable nuclides freshly produced and ejected by the explosion. Infrared, optical, and X-ray spectroscopic measurements are capable of determining elemental abundances in the photosphere during different phases of the outburst. But the results from such observations are, in general, sensitive to the models of line excitations in the photosphere, resulting in large correction factors to be applied before deducing the isotopic yields at the time of the explosion. In addition, results are hampered by the uncertainty of the optical depth, and by the possibility of heavy element condensation into dust shortly after the explosion, as was witnessed in SN 1987A. A more direct probe of massive-star interior physics is, in principle, to investigate unstable nuclides and to measure the $\gamma$-rays associated with their $\beta$ decays after they have been ejected by the supernova explosion. For an optimum probe, the mean lifetime of such a radioactive isotope should range from around a few weeks up to about 10$^6$ years. The lower limit is set by the requirement that the ejecta should become optically thin to $\gamma$-rays in a few decay times, and the upper limit by instrumental sensitivities of $\gamma$-ray telescopes (the $\gamma$-ray flux from a trace isotope must exceed the instrumental noise level, presently in the range of 10$^{-5}$ photons cm$^{-2}$s$^{-1}$, which corresponds, for instance, to (several times) $10^{-2}$ M$_\odot$ of an intermediate-mass isotope with a lifetime of $10^6$ y at the Galactic center or to $10^{-4}$ M$_\odot$ of the isotope in a supernova at a distance of a few hundreds of pc). Furthermore, whereas short-lived isotopes will clearly trace individual events, long-lived ones with mean lifetimes of the order of 10$^6$ y or longer will reflect a superposition of different supernovae at different times, mixed with interstellar matter. Consequently clues to abundances in an individual object are only very indirect. Only a few isotopes fulfill those constraints (see e.g. Diehl and Timmes 1998). Most promising cases are found among the Fe group elements, primarily because of their expected large abundances. The 0.847 and 1.238 MeV $\gamma$-ray lines from the $^{56}$Co $\rightarrow ^{56}$Fe decay (half-life: $t_{1/2}$\ = 77 d) were detected from SN 1987A (Matz et al. 1988; Sandie et al. 1988; Mahoney et al. 1988; Rester et al. 1988; Teegarden et al. 1989). There is also evidence for these decay lines from the unusually fast and bright Type Ia supernova 1991T (Morris et al. 1995, 1997). The $^{57}$Co $\rightarrow ^{57}$Fe decay ($t_{1/2}$\ = 272 d) is another probe: the 122 and 136 keV lines were detected from SN 1987A (Kurfess et al. 1992; Clayton et al. 1992). Cases at the upper end of the favored radioactive lifetime range are $^{26}$Al ($t_{1/2}$\ = 7.4 $\times 10^5$ y) and $^{60}$Fe ($t_{1/2}$\ = 1.5 $\times 10^6$ y). The 1.809 MeV line from the $^{26}$Al decay has been detected and mapped along the entire plane of the Galaxy (see review by Prantzos \& Diehl 1996). If supernovae, rather than Wolf Rayet stars, were responsible for this $^{26}$Al, the lines from the $^{60}$Fe $\beta^-$ decay would be expected simultaneously with the $^{26}$Al decay, identifying a supernova origin (review by Diehl and Timmes 1998). Instrumental sensitivity appears just at the borderline for this test. In this paper, we focus our discussion on $^{44}$Ti, which decays with $t_{1/2}$\ = 60 y (Ahmad et al. 1998; G\"orres et al. 1998; Norman et al. 1998; Wietfeldt et al. 1999), making it ideal for a study of inner-supernova physics within young supernova remnants. $^{44}$Ti decays almost uniquely to the 2$^{nd}$ excited state of $^{44}$Sc, followed immediately by the almost unique $\beta^+$ decay of $^{44}$Sc ($t_{1/2}$\ = 4 h) to the 1.156 MeV excited state of $^{44}$Ca. The 1.156 MeV de-excitation line has indeed been observed by the COMPTEL telescope on the Compton Observatory from Cas~A, a young supernova remnant with an estimated age of 320 y (Iyudin et al. 1994, 1997; Dupraz et al. 1997). The measured $\gamma$-ray flux is $\simeq(4 \pm 1)\times 10^{-5}$ photons /cm$^2$/s (Iyudin et al. 1997) concordant with an upper limit obtained by the OSSE instrument (The et al. 1996). With an adopted distance to Cas~A of 3.4 kpc (Reed et al. 1995) and the laboratory decay rate, the inferred initial mass of $^{44}$Ti is $\simeq$ $2 \times 10^{-4}$ M$_{\odot}$ (Iyudin et al. 1997; Woosley \& Diehl 1998). The current model predictions of the $^{44}$Ti\ initial mass lie in an approximate range of $(6 \times 10^{-6} \sim 8 \times 10^{-5})$ M$_\odot$ for Type-II SNe (Woosley \& Weaver 1995; Thielemann et al. 1996), and of $(3 \sim 8) \times 10^{-5}$ M$_\odot$ for Type-Ib SNe (Woosley et al. 1995), more or less strongly depending on the progenitor masses. Higher values up to $2 \times 10^{-4}$ M$_\odot$ were obtained in some of the Type-II SN models, when progenitor masses above 30 M$_\odot$ combine with high explosion energies (Woosley \& Weaver 1995), and for a 20 M$_\odot$ (SN1987A) model star (Thielemann et al.~1996). Barring the possibility that the progenitor of Cas~A happened to be such a star, one may conclude that COMPTEL observed significantly more $^{44}$Ti\ than expected (see, e.g., Fesen \& Becker 1991; Hurford \& Fesen 1996 for discussions on the progenitor characteristics). As far as its $\beta$-decay properties are concerned, $^{44}$Ti\ is a very interesting trace isotope, because its decay mode is pure orbital electron capture, which means that fully ionized $^{44}$Ti\ is stable. Even partial ionization of the innermost electrons should lead to a considerably longer effective half-life (Mochizuki 1999). Therefore, the question arises whether in supernova remnants $^{44}$Ti\ could be highly ionized and thus more stable for a considerable period of time during the evolution. In this case, it would be incorrect to use the half-life measured in the laboratory, and initial abundances of $^{44}$Ti\ as deduced from $\gamma$-ray intensities could be too high. However, there is no simple answer to this question and, as we shall emphasize below, the thermodynamic history of a remnant has to be known in detail before firm predictions can be made. On the other hand, it is interesting to speculate whether the COMPTEL observations, which indicate an amount of $^{44}$Ti\ in Cas~A that appears higher than expected, reflect the effects of an increased lifetime of the explosively-produced $^{44}$Ti\ because of temporary and partial ionization. The primary aim of the present paper is to outline possible implications of (partial) ionization on the observable $\gamma$-ray line flux from the decay of $^{44}$Ti . We shall present results obtained for a variety of different conditions, based on a simple model for young supernova remnants which, nonetheless, accounts for the features most relevant to this question. Of course, our main focus will be on Cas~A, the best studied case, and the parameters of the model are chosen accordingly. Section 2 contains an overview of observational aspects of SN explosions and remnants which are of relevance in the context of this work. In Sect.~3, we describe the employed model for young supernova remnants, i.e., the analytic model of McKee \& Truelove (1995), augmented by a description of the reverse shock interacting with denser cloudlets (Sgro 1973; Miyata 1996). In addition, we describe the microphysics used in the model on which the calculation of the ``effective'' decay rate of $^{44}$Ti\ is based. In Sect.~4, we present our results for the time variation of the $^{44}$Ti\ decay rate, the $^{44}$Ti\ abundance in young supernova remnants and the associated $\gamma$-ray activities that can be measured. As a specific example, we will preferentially use the case of Cas~A. Summary and conclusions are given in Sect.~5. \section{Young Supernova Remnants} Studies of supernovae from the moment of the explosion up until they dissolve and merge with the general interstellar medium has, in general, been guided by observational opportunities as combined with model interpretations; the interplay of different physical processes, which vary spatially and rapidly within a young supernova remnant, cannot be disentangled from measurements alone. Observational windows are (from low to high energy radiation): Radio maps from electrons synchrotron-emitting in magnetic-field structures around the shocked-gas region; infrared emission from cool and hot dust within dense clumps embedded in the supernova remnant; optical line emission from the edges of dense material embedded in the remnant's hot plasma; X-ray emission from the hot, ionized, gas that has been shocked by the forward blast wave and by the reverse shock traveling inward, respectively; and $\gamma$-rays from long-lived radioactivity such as $^{44}$Ti . Here we are mainly interested in the thermal history of the radioactive $^{44}$Ti\ after it has been ejected in a supernova explosion. $^{44}$Ti\ as well as other iron-group elements are synthesized during the very early moments of the explosion of a massive star in the layers adjacent to the nascent compact object (a neutron star or black hole). Therefore, observations of iron can be used to trace also $^{44}$Ti . However, not many good cases are known so far. SN~1987A is certainly the best studied case. The large Doppler shifts of iron lines observed in the early spectra require that iron moves with velocities much higher than some of the hydrogen which cannot be explained by spherically symmetric explosion models but indicates that it is not homogeneously distributed in an expanding shell but is found in clumpy structures. This conclusion is also supported by the unexpectedly early detection of X- and $\gamma$-rays from the decay of radioactive elements (for summaries, see Woosley \& Weaver 1994, Nomoto et al.~1994), which again suggests that these nuclei have been transported far out into the hydrogen-rich shells. Other arguments in favor of this interpretation include the smoothness of the light-curve (e.g., Woosley \& Weaver~1994, Nomoto et al.~1994 and references therein), and the time-dependent features of the spectral lines observed soon after the outburst (Utrobin et al. 1995), e.g., in the Bochum event (Hanuschik \& Dachs 1987; Phillips \& Heathcote 1989). In fact multi-dimensional supernova simulations (e.g., Herant et al. 1992, 1994; Shimizu et al. 1994; Shimizu 1995; Burrows et al. 1995; Janka \& M\"uller 1995, 1996) have demonstrated that the surroundings of the newly-formed neutron star are stirred by hydrodynamic instabilities and that inhomogeneities and clumpiness of the products of explosive nucleosynthesis are likely the consequence. However, it cannot be excluded that SN 1987A is a special case and not comparable with, e.g., Cas~A. $^{44}$Ti\ $\gamma$-ray emission is expected from core-collapse supernovae in general (e.g., Woosley \& Weaver 1995; Thielemann et al. 1996), yet there is a significant deficit in such $\gamma$-ray line sources in the Galaxy for the inferred Galactic core-collapse supernova rate (Dupraz et al. 1997). The $^{44}$Ti\ detection for Cas~A (Iyudin et al. 1994; 1997), despite the low Fe abundance, therefore has given rise to speculations about its exceptional nature, particularly because its optical and X-ray characteristics support the idea of asymmetries and a peculiar circumstellar environment (see below, and Hartmann et al.~1997 and references therein). It was suggested, for example, that the progenitor of Cas~A might have been a rapidly spinning star, in which case more $^{44}$Ti could have been synthesized (Nagataki et al. 1998). Recently, a second Galactic $^{44}$Ti\ source has been reported (Iyudin et al. 1998). Its alignment with an also recently discovered X-ray remnant (Aschenbach 1998) suggests that it is a very young and most nearby supernova remnant, with an age around 680y and a distance of 200pc only. This object still provides a puzzle because of the absence of radio and optical emission expected for such a nearby supernova. Yet, if confirmed, it will provide a unique opportunity for the study of supernova-produced $^{44}$Ti; in which case it is of some interest to see if the modified decay rate of ionized $^{44}$Ti\ addressed in our paper could also be important. The $^{44}$Ti\ found in young supernova remnants is probably formed during the $\alpha$-rich freezeout from (near) nuclear statistical equilibrium (Woosley et al. 1973; Woosley \& Weaver 1995; Woosley et al. 1995; Thielemann et al. 1996; Timmes et al. 1996; Timmes \& Woosley 1997; The et al. 1998). Whereas $^{44}$Ti\ is obviously fully ionized at the time of explosion, $^{44}$Ti\ ions will recombine with electrons during the adiabatic cooling phase, which accompanies the expansion of the exploded star, to become neutral after some 1000 s (e.g., Nomoto et al.~1994), a negligible timescale when compared with the age of a supernova remnant. Therefore, during most of this early phase $^{44}$Ti\ will decay with the laboratory rate. The subsequent evolution of the remnant may however provide conditions for its re-ionization, which is the subject of our modeling effort. The evolution of young supernova remnants has been modeled extensively, and the spherically-symmetric explosion into homogeneous interstellar surroundings appears well-understood (McKee and Truelove 1995). A free expansion phase is followed by an adiabatic blast-wave phase, where interaction with surrounding material produces outward and inward moving shock waves leading to bright X-ray and radio emission, yet being unimportant for the energetics of the remnant. Later phases of significant slowing-down and radiative losses of the expanding remnant lie beyond the early phase where $^{44}$Ti\ still decays. Young remnants such as Cas~A are generally understood as being somewhere intermediate between free expansion and the second phase, commonly called ``Sedov-Taylor'' phase. The evolution of such idealized young supernova remnants during the ejecta-dominated stage ($t < t_{\rm ST}$) and the Sedov-Taylor phase ($t > t_{\rm ST}$) was described in an analytical model by McKee \& Truelove (1995). At the interface between ejecta and ambient medium a contact discontinuity occurs, separating the exterior region of shocked and swept-up circumstellar gas behind the outward-moving blastwave from the supernova remnant interior. Inward from the discontinuity, a reverse shock travels through the expanding, cold supernova ejecta, heating up the interior gas. Bright X-ray emission results from the hot plasma on both sides of the contact discontinuity, with higher temperature on the outside heated by the blastwave ($\geq$ keV), as compared with reverse-shock heated gas on the inside ($\leq$ keV; e.g. Vink et al. 1996). Although such a two-temperature model appears adequate to describe the X-ray emission of many supernova remnants, considerable uncertainty remains in detail. For example, the blastwave shock rapidly heats the ions entering the blastwave region, the thermalization times for the X-ray emitting electrons may exceed 100 years, so that non-equilibrium models are needed for a proper description of X-ray and radio emission. The detailed physical conditions within supernova remnants are far from being understood, although the general evolution follows these fairly simple descriptions quite closely, and is controlled by a few parameters, the explosion energy, the mass of ejecta, and the surrounding medium density. According to the McKee \& Truelove (1995) model, an explosion energy in the $(1 \sim 3) \times 10^{51}\,{\rm erg}$ range does not seem unreasonable in the case of Cas~A for the possible ejecta mass of $(2 \sim 5)$ M$_\odot$ (Tsunemi et al. 1986; Vink et al. 1996) if the surrounding ambient gas density is of the order of 20 cm$^{-3}$ (Tsunemi et al. 1986) and the current blastwave radius is $(2 \sim 3)$ pc for an assumed distance of 3.4 kpc (Anderson \& Rudnick 1995; Holt et al. 1994; Jansen et al. 1988; Reed et al. 1995). Although the general appearance of supernova remnant images in the radio and in X-rays is that of a large-scale shell configuration as expected from the above model, additional prominent clumpy structures appear in some cases (e.g., Anderson \& Rudnick 1995; Koralesky et al. 1998). This suggests that the model outlined so far is an oversimplification as far as details are concerned. The gross radio and X-ray emissivity may not be very sensitive to such discrepancy, tracing the electron component in the vicinity of the shock region, but the bulk ejected mass may be inadequately represented by the inferred electron densities and temperatures (e.g., Koralesky et al. 1998). Thus, for the Cas~A remnant, prominent structures have been studied in their respective forms of optical knots and filaments, ``quasi-stationary flocculi", ``fast-moving knots", and ``fast-moving flocculi" (e.g., van den Bergh \& Kamper 1983; Reed et al. 1995; Peimbert \& van den Bergh 1971; Chevalier \& Kirshner 1977, 1978, 1979; Reynoso et al. 1997; Lagage et al. 1996). There is overwhelming evidence for dense structures embedded in tenuous material within the entire remnant, and even outside the blastwave shock radius. Obviously the explosion itself produces fragments of material, seen now as fast-moving knots with their abundance patterns supporting an ejecta origin. These clumps might carry heavier elements preferentially as suggested by observations of fast-moving Fe clumps early in SN explosions, e.g. in SN 1987A (e.g., Nomoto et al.~1994, Wooden~1997 and references therein), but a connection between the instabilities and clumpiness early after the supernova explosion and the fragments and ``bullets'' seen in the remnants has not been established yet. Given all these uncertainties in the evolution of supernova remnants we do not attempt to model a specific object, such as Cas~A, in detail here. We rather shall investigate by means of an admittedly very simple model, varying its parameters within reasonable limits, the potential effects of ionization on the $^{44}$Ti\ abundance estimates obtained from $\gamma$-ray observations. \section{The Model} The remnant model employed in this work uses the analytic description by McKee \& Truelove (1995) for the hydrodynamic evolution. The ejecta with mass $M_{\rm ej}$ are assumed to be cold and to expand with explosion energy $E_{\rm ej}$ into a homogeneous ambient medium which has a hydrogen number density $n_{\rm H0}$ and is composed of 10 hydrogen atoms per helium atom, corresponding to a helium mass fraction of 28.6\%. Inside the homogeneous ejecta, we assume the existence of denser clumps that contain a small fraction of the ejecta mass but most of the produced $^{44}$Ti. The treatment of the interaction of the reverse shock with these clumps and of the corresponding effects on the $^{44}$Ti\ decay are described below. Ionization of $^{44}$Ti\ is most efficiently induced by electrons. If and how long a high degree of ionization is sustained depends on the reverse-shock characteristics and the postshock evolution of the matter containing $^{44}$Ti. \subsection{A model for young supernova remnants} If sufficiently high temperatures happen to be reached behind the reverse shock, $^{44}$Ti may become fully ionized, in which case its decay is prevented until cooling due to the expansion of the gas leads to (partial) recombination of the innermost electrons. To follow the radioactivity of the supernova remnant requires knowledge of the evolution of the shocked ejecta. The analytical model of McKee \& Truelove (1995) yields scaled relations for the radius, velocity and postshock temperature of the reverse shock as functions of time $t$. Also, the model provides the density of the unshocked (homogeneous) ejecta at time $t$, which allows one to calculate the postshock density from the density jump at the reverse shock. In order to describe the density history of each mass shell after the reverse shock has passed through it, we assume that the shocked matter moves (approximately) with the same velocity as the contact discontinuity. Knowing the motion of the latter, therefore, we can estimate the dilution of the shocked gas by the expansion of the supernova ejecta. The time-dependent analytical solutions are given by the three parameters, $M_{\rm ej}$, $E_{\rm ej}$ and $n_{\rm H0}$. A combination of values of these parameters is considered to yield a suitable description of a certain supernova remnant, if the model gives a present-day radius and velocity of the blastwave that is compatible with observational data. We extended the analytic remnant model of McKee \& Truelove (1995) by assuming that the iron-group elements together with the explosively nucleosynthesized $^{44}$Ti\ are concentrated in overdense clumps of gas. These clumps are assumed to move outward through the dominant mass of homogeneous supernova ejecta with high velocity. Since the clumps should contain only a minor fraction of the total mass of the ejecta, the dynamics of the young supernova remnant as described by the McKee \& Truelove (1995) model is assumed not to be altered by their presence. The radial position $r$ of the iron clumps within the ejecta can be measured by the relative mass coordinate $q$ which is defined as \begin{equation} q \equiv\, {4\pi r^3\rho_{\rm ej}\over 3\,M_{\rm ej}}\, , \end{equation} \noindent where $\rho_{\rm ej}$ is the uniform background density of the ejecta which expand homogeneously. \subsection{Thermodynamic conditions just behind the reverse shock} We now consider the properties of a clump of metal-rich matter within the homogeneous ejecta. Denoting the density enhancement factor of the assumed clump relative to the surrounding ejecta by $\alpha_{\rm clmp}$, we obtain the clump density after the reverse shock has passed at time $t_{\rm sh}$ as \begin{equation} \rho (t_{\rm sh}) = 4\, \alpha_{\rm clmp}\, \rho_{\rm ej} (t_{\rm sh}) \, , \end{equation} \noindent for an adiabatic index $\gamma = 5/3$. Correspondingly, the post-reverse shock temperature of the clump is related to that of the surrounding uniform ejecta, $T_{\rm ej}(t_{\rm sh})$, as \begin{equation} T (t_{\rm sh}) = \frac{\beta(\alpha_{\rm clmp})}{\alpha_{\rm clmp}}\, \frac{\mu(t_{\rm sh})}{\mu_{\rm ej}}\, T_{\rm ej}(t_{\rm sh})\, , \end{equation} \noindent which generalizes by the factor $\mu(t_{\rm sh})/\mu_{\rm ej}$ the result of Sgro (1975) to the case considered here, where the chemical composition of the clump is different from that of the surrounding ejecta as described by the corresponding mean molecular weights $\mu(t_{\rm sh})$ and $\mu_{\rm ej}$, respectively. $T_{\rm ej}(t_{\rm sh})$ is given by \begin{equation} T_{\rm ej}(t_{\rm sh}) = \frac{3 \mu_{\rm ej}}{16 k}\, {\widetilde v}_{\rm r}^{\,2}(t_{\rm sh})\, , \end{equation} where ${\widetilde v}_{\rm r}$ is the reverse-shock velocity in the rest frame of the unshocked ejecta (McKee \& Truelove 1995). In Eq.~(3), $\beta$ is the ratio of the pressure values in the regions behind the reflected shock and in front of it. The reflected shock is formed when the reverse shock hits a dense clump in the ejecta (Sgro 1975; Miyata 1996). The ratio $\beta$ is related to $\alpha_{\rm clmp}$ as \begin{equation} \alpha_{\rm clmp} = \beta\, \Bigl( 1 + \frac{1-\beta}{\sqrt{4 \beta +1}} \Bigr)^{-2}\, , \end{equation} \noindent for $\gamma = 5/3$ (Sgro 1975). In particular, $\beta = 1$ if $\alpha_{\rm clmp} = 1$ (no clumpiness), and $\beta/\alpha_{\rm clmp}$ appearing in Eq.~(3) is less than unity for $\alpha_{\rm clmp} > 1.$ \subsection{Post reverse-shock evolution} We now follow the fate of the homogeneous ejecta and of the matter in clumps after they have experienced the impact of the the reverse shock at time $t_{\rm sh}$. As we have mentioned earlier, we assume that the shock itself would continue to proceed inward through the cold ejecta as prescribed by the McKee \& Truelove (1995) model. The thermodynamic and chemical evolution of the clumps at times $t \geq t_{\rm sh}$\ is treated as described below. Since we trace the clump material, it is preferable to introduce Lagrangian particle abundances, defined by \begin{equation} n \equiv \frac{{\rm d} N }{{\rm d} q}\, , \end{equation} \noindent where d$N$ is the number of particles at the mass coordinate $q$ in the interval d$q$. \subsubsection{Ionic abundances} At a given time $t (\geq t_{\rm sh})$, the number abundance $n_i$ of a nuclear species $k$ with charge $Z$ and $i$ orbital electrons (i.e. $Z - i$ times ionized, and $n_k = \sum_{i \geq 0} n_i)$ follows \begin{eqnarray} {\frac{{\rm d} n_i}{{\rm d} t}} = - [\lambda_{\beta,i} + \lambda_{{\rm ion},i} + \lambda_{{\rm rec},i}] n_i \nonumber \\ +\lambda_{{\rm ion},i+1} n_{i+1} + \lambda_{{\rm rec},i-1} n_{i-1}\, , \end{eqnarray} \noindent where $\lambda_{\beta,i}(t), \lambda_{{\rm ion},i}(t),$ and $\lambda_{{\rm rec},i}(t)$ are the nuclear $\beta$-decay, ionization and recombination rates, respectively, of the ion. For $^{44}$Ti, $\lambda_\beta \neq 0$, and thus $n(t>t_{\rm sh}) < n(t_{\rm sh}) = n(0) {\rm exp}( - \lambda_{\rm lab} t_{\rm sh} ),$ where $\lambda_{\rm lab}$ is the $^{44}$Ti\ decay rate in the cold ejecta, i.e. its laboratory value. Since we assume that material which contains $^{44}$Ti\ consists dominantly of one single stable nuclide ($^{56}$Fe in practice), the number of ionization electrons, $n_{\rm e}$, is given by $\sum_{i \geq 0} (Z - i) n_i$ for that species. In practice, we further assume that both $^{56}$Fe and $^{44}$Ti\ are ionized to some degree in the reverse shock itself, which we cannot resolve. Typically, we start our network calculations with ionic systems of 10 bound electrons, This choice is made since our major concern is to see the reduction of the orbital-electron capture rates in the domains of high shock temperatures. The reaction rates entering in Eq.~(7) are discussed in Appendix A. \subsubsection{Ion and electron temperatures} The network requires the knowledge of the thermodynamic evolution of the shocked matter, and in particular of the electron temperature, $T_{\rm e}$, which defines the Boltzmann distribution, and of the matter density $\rho$ (or the ion number density, $n_{\rm ion}$), which enters in the absolute electron number density $n_{\rm e}$. The decrease by expansion of the temperature $T(t)$ from $T(t_{\rm sh})$ of Eq.~(3) would roughly be proportional to $[\rho(t)/\rho(t_{\rm sh})]^{2/3}$ (for $\gamma = 3/5$ and $t > t_{\rm sh})$. Two considerations are due, however. One concerns the possibility that both $T_{\rm e}$ and the ion temperature, $T_{\rm ion}$, may not necessarily be the same as $T$. Secondly, some non-adiabatic effects may appear as the result of ionization and recombination. It is sometimes asserted that the collisionless equilibration between $T_{\rm ion}$ and $T_{\rm e}$ would occur quickly, in which case $T_{\rm ion} = T_{\rm e} = T$ already at $t = t_{\rm sh}$. Save this possibility, the equipartition of the shock energy to ions and electrons may not be achieved at once. The equilibration process through the Coulomb interaction is described by a non-linear equation (Spitzer 1962), \begin{equation} \frac{{\rm d} T_{\rm e}}{{\rm d} t} = \frac{T_{\rm ion} - T_{\rm e}}{\tau_{\rm eq}}\, , \end{equation} \noindent in terms of the equilibration time $\tau_{\rm eq}(t)$, which is a function primarily of $T_{\rm ion}$, $T_{\rm e}$ and the number densities of ions and electrons. Assuming that $T_{\rm e} \ll T_{\rm ion}$ at $t = t_{\rm sh}$, and that the total kinetic energy \begin{equation} \frac{3}{2} ( n_{\rm ion} + n_{\rm e} ) T = \frac{3}{2} ( n_{\rm ion} T_{\rm ion} + n_{\rm e} T_{\rm e} ) \end{equation} \noindent is constant during a short time interval $\delta t$, we solve Eq.~(8) simultaneously with the abundance network equations to determine the degree of ionization of a dominant species, and thus $n_{\rm e}$. In order to include the non-adiabatic effect caused by ionization and recombination, we assume that the energy required for unbinding additional electrons (net energy loss) is taken from electrons only. Namely, after performing the network calculation at each time step and solving Eq.~(8), we make a replacement \begin{equation} T_{\rm e} \rightarrow T_{\rm e} {\rm exp}\, \Bigl( - \frac{\delta n_{\rm e}}{n_{\rm e}} - \frac{2}{3} \frac{ \delta Q}{n_{\rm e} T_{\rm e}}\Bigr)\, , \end{equation} \noindent where $\delta n_{\rm e}$ and $\delta Q$ are the change of $n_{\rm e}$ during $\delta t$ and the corresponding net energy consumption. The equilibrium temperature $T$ is then calculated from Eq.~(9). Finally, $T(t + \delta t), T_{\rm ion}(t + \delta t),$ and $T_{\rm e}(t + \delta t)$ are obtained from the current values by multiplying $[\rho(t +\delta t) / \rho(t)]^{2/3}.$ \section{Results} We present here our results, which we obtained with the model described above. They depend on the following parameters: $E_{\rm ej}, M_{\rm ej}, n_{\rm H0}$, $\alpha_{\rm clmp}$ and two values for $q$, given as $q_0$ and $q_1$, which define the lower and upper boundaries of the mass shell where the $^{44}$Ti-containing clumps are assumed to exist. As we mentioned earlier, the values of the first three can be constrained in the case of Cas~A from the observational analyses of the blast-wave radius $R_{\rm b}$ and velocity $v_{\rm b}$. Below we consider mainly the parameter space which is consistent with these constraints. While we treat $\alpha_{\rm clmp}$ as a free parameter, a value of the order of 10 seems to be reasonable as a first guess, based on the density contrast which develops in hydrodynamic instabilities in the envelope of the progenitor star during the supernova explosion (Fryxell et al. 1991). \begin{figure} \epsfxsize=0.99\hsize \epsffile{ti44fig01.ps} \caption{Examples for the evolution with time $t$ (in years) of the $^{56}$Fe number density $n_{\rm Fe}$ ({\it top}), and of the electron temperature $T_{\rm e}$ ({\it bottom}) in the $^{56}$Fe-dominated clumps at four different locations indicated by the mass coordinate $q$. The dashed lines represent the asymptotic behavior of the equilibrium temperatures $T(t)$ (as described in Sect.~3.3.2) at $q = 0.4$ and $0.6$. The left edge of each solid curve corresponds to $t = t_{\rm sh}$. For example, the reverse shock took 50 y to reach $q = 0.6$. The results are given for the following set of parameter values: $E_{\rm ej} = 3 \times 10^{51}$ erg, $M_{\rm ej} = 3$ M$_\odot, n_{\rm H0} = 15$ cm$^{-3}$ and $\alpha_{\rm clmp} = 10$ } \end{figure} \subsection{Thermodynamic evolution of post-shock clumps} Figure 1 shows the evolution of the $^{56}$Fe number density $n_{\rm Fe}$ and of the electron temperature $T_{\rm e}$ in the $^{56}$Fe-dominated clumps at four different locations in the ejecta, given by the corresponding mass coordinate $q$, after the reverse shock has passed the clumps. The post-shock expansion decreases the density as well as the equilibrium temperature, while $T_{\rm e}$ first increases in its attempt to reach equilibrium between electrons and ions. The results are plotted for the chosen set of parameter values quoted therewith. \begin{figure} \epsfxsize=0.99\hsize \epsffile{ti44fig02ab.ps} \epsfxsize=0.99\hsize \epsffile{ti44fig02c.ps} \caption{{\it Top panel:} The average numbers $N_{\rm e}$ of electrons that are bound to $^{44}$Ti\ and to $^{56}$Fe at mass coordinate $q$ and at time $t$ (in y) after the explosion. {\it Middle:} Effective decay rates $\Lambda_{\rm postsh}$ of the shocked $^{44}$Ti\ during the time span between $t_{\rm sh}$\ and $t$ [Eq.~(11)] in units of the laboratory rate $\lambda_{\rm lab}$. {\it Bottom:} The corresponding $^{44}$Ti\ radioactivity observable at time $t$ (``age'') relative to the case that assumes no reduction of the $\beta$-decay rate. The parameter values used here are the same as those for Fig.~1 } \end{figure} \begin{figure} \epsfxsize=0.99\hsize \epsffile{ti44fig03.ps} \caption{$^{44}$Ti\ radioactivities at times $t$ (in y) in the clumps located at radial positions $q$, relative to the radioactivity of the model with most detailed input physics, when the non-adiabaticity of the thermal evolution of the clumps is ignored ({\it upper panel}) and when instantaneous equilibration between $T_{\rm e}$ and $T_{\rm ion}$ (toward $T$) is assumed ({\it lower panel}) } \end{figure} \subsection{Effective decay rates of $^{44}${\rm Ti} and its radioactivity as an observable} Let us now consider $^{44}$Ti\ in a clump that is located anywhere in the mass coordinate $0 \leq q \leq 1$ of the ejecta. If $^{44}$Ti\ is near the surface (large $q$), it suffers from the reverse shock early. Since the electron temperature is still relatively low (see Fig.~1), there is no possibility to reach a high degree of ionization (thus an appreciable reduction of the decay rate). On the other hand, $^{44}$Ti\ embedded in clumps located in the innermost region may become highly ionized as the temperature is generally highest, resulting in the smallest $^{44}$Ti\ decay rate. Since $t_{\rm sh}$ is closer to $t$, the ``age'' of the remnant, however, the net effect of the reduced decay rate appears weaker. Therefore, the maximum effect of the retarded $\beta$-decay is obtained if $^{44}$Ti\ is at intermediate values of $q$. This situation is depicted in the panels of Fig.~2. The top panel of Fig.~2 shows the average numbers of electrons that are bound to $^{44}$Ti\ and to $^{56}$Fe at $t = 100, 200$ and 300 years after the explosion. The increase of $N_{\rm e}$ toward the low side of $q$ indicates that the time span between $t_{\rm sh}$ and $t$ was not long enough for ionization. At the high end of $q$-values, the effects of post-shock recombination are visible. When combined with $n_{\rm Fe}$ from Fig.~1, $N_{\rm e}(^{56}$Fe) determines the total number of ionization electrons, $n_{\rm e}(t)$, in a clump: $n_{\rm e} = [Z - N_{\rm e}(^{56}$Fe)] $n_{\rm Fe}$ with $Z = 26$. The retardation of the $^{44}$Ti\ decay by ionization is illustrated in the middle panel of Fig.~2. The effective decay rate in the post-shock period, $\Lambda_{\rm postsh}$, is defined as the time-average of $\lambda_{\rm eff}$ [given by Eq.~(18) in Appendix A] between the time $t_{\rm sh}$\ and a time $t > t_{\rm sh}$. It can be expressed as \begin{equation} \Lambda_{\rm postsh} \equiv - \frac{1}{t - t_{\rm sh}}\, {\rm ln}\Bigl[ \frac{n(t)}{n(t_{\rm sh})} \Bigr]\, , \end{equation} \noindent where $n(t)$ is the $^{44}$Ti\ abundance at mass coordinate $q$ and time (age) $t$, and $n(t_{\rm sh})$ is the corresponding value at the shock impact time $t_{\rm sh}$, which has been reduced from its initial value $n(0)$ according to $n(t_{\rm sh}) = n(0) {\rm exp}(- \lambda_{\rm lab} t_{\rm sh})$. In order to estimate the observable consequences of the reduced $\beta$-decay rates, one has to recall that the measurable $\gamma$-ray activity per (normalized) unit mass of the remnant is the product of the current $\beta$-decay rate and the current abundance, $n(t) \lambda_{\rm eff}(t)$. This introduces a nonlinear relation between the observable activity given in the bottom panel of Fig.~2 and the effective decay rates displayed above, particularly in the absence of quick recombination. In Fig.~3, we compare the results obtained with the input physics as described in Sect.~3 to the cases where either the non-adiabaticity of the equilibration process of electrons and ions owing to the energy consumption by the ionization process is neglected ({\it upper panel}), or where instantaneous equilibration (i.e., $T_{\rm e} = T_{\rm ion} = T$ always) is assumed ({\it lower panel}). It can be seen that in both cases the trends of the ``full'' model shown in Fig.~2, where both the non-adiabaticity and the gradual equilibration process are taken into account, are somewhat enhanced. This can be understood by the fact that the omission of either of the two effects leads to higher electron temperatures and thus to a higher degree of ionization in the clumps. \subsection{$^{44}${\rm Ti} radioactivity in Cas~A} The $^{44}$Ti\ radioactivity of the whole supernova remnant as observable at an age $t$, relative to the activity level that would result in the absence of any retardation of the $\beta$-decay rate by ionization, is defined by the ratio \begin{equation} {\cal A} \equiv \int_{q_0}^{q_1} n(q, t) \lambda_{\rm eff}(q, t) {\rm d}q {\bigg /} \Bigl[ N(0){\rm exp}( - \lambda_{\rm lab} t) \lambda_{\rm lab} \Bigr]\, . \end{equation} \noindent Here, $N(0)$ is the total number of $^{44}$Ti\ nuclei in the $^{56}$Fe clumps at $t = 0$. For a remnant like Cas~A, which is in the transition phase from the ejecta-dominated stage to the Sedov-Taylor phase, the reverse shock has passed through most of the ejecta. Therefore, the factor ${\cal A}$ must be expected to be larger than unity when the remnant is observed at this stage of its evolution. This implies that the initial abundance inferred from the current $\gamma$-ray activity due to the $^{44}$Ti\ decay is lower by a factor ${\cal A}^{-1}$ than the amount of produced $^{44}$Ti\ that is estimated on the basis of the laboratory decay rate $\lambda_{\rm lab}$. \begin{table} \caption{Illustration of the variability of the $^{44}$Ti\ radioactivity ratios ${\cal A}$ as defined by Eq.~(12) with different values of the model parameters. The tabulated results are at the time (age) of 320 y after the explosion for an explosion energy of $E_{\rm ej} = 1 \times 10^{51}$ erg and various combinations of values for the ejecta mass $M_{\rm ej}$, the hydrogen number density in the ambient medium, $n_{\rm H0}$, the over-density factor of the clumps, $\alpha_{\rm clmp}$, and the lower and the upper boundaries $q_0$ and $q_1$ of the mass shell where the clumps are assumed to exist. $R_{\rm b}$ and $v_{\rm b}$ are the radius and velocity of the blast-wave at that time as computed from the McKee \& Truelove (1995) model for the given sets of $E_{\rm ej}, M_{\rm ej}$ and $n_{\rm H0}$ values } $$\vbox{\tabskip4pt \halign{#&#&#&#&#&#&#&# \cr \noalign{\hrule} \noalign{\smallskip} \hfil $M_{\rm ej}$ & \, $n_{\rm H0}$ & \ $R_{\rm b}$ & \ $v_{\rm b}$ & \, $\alpha_{\rm clmp}$ & \ $q_0$ & \ $q_1$ & \ ${\cal A}$ \cr (M$_\odot$) & (cm$^{-3}$) & (pc) & $(10^3$ km/s) & & & \cr \noalign{\smallskip} \noalign{\hrule} \noalign{\smallskip} \noalign{\smallskip} \ 2 & \ 15 & 1.71 & \ 2.53 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.47 \cr \ 2 & \ 15 & 1.71 & \ 2.53 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.12 \cr \ 2 & \ 15 & 1.71 & \ 2.53 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.55 \cr \ 2 & \ 15 & 1.71 & \ 2.53 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.27 \cr \ 2 & \ 15 & 1.71 & \ 2.53 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.34 \cr \noalign{\smallskip} \ 2 & \ 30 & 1.51 & \ 2.15 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.92 \cr \ 2 & \ 30 & 1.51 & \ 2.15 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.30 \cr \ 2 & \ 30 & 1.51 & \ 2.15 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.77 \cr \ 2 & \ 30 & 1.51 & \ 2.15 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.43 \cr \ 2 & \ 30 & 1.51 & \ 2.15 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.35 \cr \noalign{\smallskip} \ 3 & \ 15 & 1.65 & \ 2.66 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.27 \cr \ 3 & \ 15 & 1.65 & \ 2.66 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.04 \cr \ 3 & \ 15 & 1.65 & \ 2.66 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.27 \cr \ 3 & \ 15 & 1.65 & \ 2.66 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.13 \cr \ 3 & \ 15 & 1.65 & \ 2.66 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.18 \cr \noalign{\smallskip} \ 3 & \ 30 & 1.47 & \ 2.23 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.55 \cr \ 3 & \ 30 & 1.47 & \ 2.23 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.13 \cr \ 3 & \ 30 & 1.47 & \ 2.23 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.38 \cr \ 3 & \ 30 & 1.47 & \ 2.23 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.22 \cr \ 3 & \ 30 & 1.47 & \ 2.23 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.20 \cr \noalign{\smallskip} \ 4 & \ 15 & 1.59 & \ 2.80 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.14 \cr \ 4 & \ 15 & 1.59 & \ 2.80 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.00 \cr \ 4 & \ 15 & 1.59 & \ 2.80 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.15 \cr \ 4 & \ 15 & 1.59 & \ 2.80 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.06 \cr \ 4 & \ 15 & 1.59 & \ 2.80 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.11 \cr \noalign{\smallskip} \ 4 & \ 30 & 1.43 & \ 2.32 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.33 \cr \ 4 & \ 30 & 1.43 & \ 2.32 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.05 \cr \ 4 & \ 30 & 1.43 & \ 2.32 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.22 \cr \ 4 & \ 30 & 1.43 & \ 2.32 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.12 \cr \ 4 & \ 30 & 1.43 & \ 2.32 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.15 \cr \noalign{\smallskip} \ 5 & \ 15 & 1.54 & \ 2.95 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.07 \cr \ 5 & \ 15 & 1.54 & \ 2.95 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 0.99 \cr \ 5 & \ 15 & 1.54 & \ 2.95 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.08 \cr \ 5 & \ 15 & 1.54 & \ 2.95 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.03 \cr \ 5 & \ 15 & 1.54 & \ 2.95 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.07 \cr \noalign{\smallskip} \ 5 & \ 30 & 1.40 & \ 2.41 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.20 \cr \ 5 & \ 30 & 1.40 & \ 2.41 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.01 \cr \ 5 & \ 30 & 1.40 & \ 2.41 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.14 \cr \ 5 & \ 30 & 1.40 & \ 2.41 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.07 \cr \ 5 & \ 30 & 1.40 & \ 2.41 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.11 \cr \noalign{\smallskip} \noalign{\hrule} }}$$ \end{table} \begin{table} \caption{Same as Table 1 but for $E_{\rm ej} = 2 \times 10^{51}$ erg } $$\vbox{\tabskip4pt \halign{#&#&#&#&#&#&#&# \cr \noalign{\hrule} \noalign{\smallskip} \hfil $M_{\rm ej}$ & \, $n_{\rm H0}$ & \ $R_{\rm b}$ & \ $v_{\rm b}$ & \, $\alpha_{\rm clmp}$ & \ $q_0$ & \ $q_1$ & \ ${\cal A}$ \cr (M$_\odot$) & (cm$^{-3}$) & (pc) & $(10^3$ km/s) & & & \cr \noalign{\smallskip} \noalign{\hrule} \noalign{\smallskip} \noalign{\smallskip} \ 2 & \ 5 & 2.44 & \ 3.63 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.19 \cr \ 2 & \ 5 & 2.44 & \ 3.63 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.03 \cr \ 2 & \ 5 & 2.44 & \ 3.63 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.33 \cr \ 2 & \ 5 & 2.44 & \ 3.63 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.26 \cr \ 2 & \ 5 & 2.44 & \ 3.63 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.51 \cr \noalign{\smallskip} \ 2 & \ 15 & 2.01 & \ 2.80 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.53 \cr \ 2 & \ 15 & 2.01 & \ 2.80 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.14 \cr \ 2 & \ 15 & 2.01 & \ 2.80 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.07 \cr \ 2 & \ 15 & 2.01 & \ 2.80 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.53 \cr \ 2 & \ 15 & 2.01 & \ 2.80 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 2.03 \cr \noalign{\smallskip} \ 2 & \ 30 & 1.77 & \ 2.40 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 2.12 \cr \ 2 & \ 30 & 1.77 & \ 2.40 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.36 \cr \ 2 & \ 30 & 1.77 & \ 2.40 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.67 \cr \ 2 & \ 30 & 1.77 & \ 2.40 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.76 \cr \ 2 & \ 30 & 1.77 & \ 2.40 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 2.11 \cr \noalign{\smallskip} \ 3 & \ 15 & 1.96 & \ 2.90 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.47 \cr \ 3 & \ 15 & 1.96 & \ 2.90 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.09 \cr \ 3 & \ 15 & 1.96 & \ 2.90 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.74 \cr \ 3 & \ 15 & 1.96 & \ 2.90 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.33 \cr \ 3 & \ 15 & 1.96 & \ 2.90 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.54 \cr \noalign{\smallskip} \ 3 & \ 30 & 1.74 & \ 2.46 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 2.04 \cr \ 3 & \ 30 & 1.74 & \ 2.46 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.30 \cr \ 3 & \ 30 & 1.74 & \ 2.46 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.13 \cr \ 3 & \ 30 & 1.74 & \ 2.46 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.53 \cr \ 3 & \ 30 & 1.74 & \ 2.46 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.57 \cr \noalign{\smallskip} \ 4 & \ 15 & 1.92 & \ 3.00 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.36 \cr \ 4 & \ 15 & 1.92 & \ 3.00 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.04 \cr \ 4 & \ 15 & 1.92 & \ 3.00 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.48 \cr \ 4 & \ 15 & 1.92 & \ 3.00 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.20 \cr \ 4 & \ 15 & 1.92 & \ 3.00 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.31 \cr \noalign{\smallskip} \ 4 & \ 30 & 1.71 & \ 2.53 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.82 \cr \ 4 & \ 30 & 1.71 & \ 2.53 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.20 \cr \ 4 & \ 30 & 1.71 & \ 2.53 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.74 \cr \ 4 & \ 30 & 1.71 & \ 2.53 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.36 \cr \ 4 & \ 30 & 1.71 & \ 2.53 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.35 \cr \noalign{\smallskip} \ 5 & \ 15 & 1.87 & \ 3.10 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.25 \cr \ 5 & \ 15 & 1.87 & \ 3.10 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.01 \cr \ 5 & \ 15 & 1.87 & \ 3.10 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.31 \cr \ 5 & \ 15 & 1.87 & \ 3.10 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.12 \cr \ 5 & \ 15 & 1.87 & \ 3.10 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.20 \cr \noalign{\smallskip} \ 5 & \ 30 & 1.68 & \ 2.59 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.61 \cr \ 5 & \ 30 & 1.68 & \ 2.59 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.11 \cr \ 5 & \ 30 & 1.68 & \ 2.59 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.49 \cr \ 5 & \ 30 & 1.68 & \ 2.59 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.23 \cr \ 5 & \ 30 & 1.68 & \ 2.59 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.24 \cr \noalign{\smallskip} \noalign{\hrule} }}$$ \end{table} \begin{table} \caption{Same as Table 1 but for $E_{\rm ej} = 3 \times 10^{51}$ erg } $$\vbox{\tabskip4pt \halign{#&#&#&#&#&#&#&# \cr \noalign{\hrule} \noalign{\smallskip} \hfil $M_{\rm ej}$ & \, $n_{\rm H0}$ & \ $R_{\rm b}$ & \ $v_{\rm b}$ & \, $\alpha_{\rm clmp}$ & \ $q_0$ & \ $q_1$ & \ ${\cal A}$ \cr (M$_\odot$) & (cm$^{-3}$) & (pc) & $(10^3$ km/s) & & & \cr \noalign{\smallskip} \noalign{\hrule} \noalign{\smallskip} \noalign{\smallskip} \ 2 & \ 5 & 2.68 & \ 3.84 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.22 \cr \ 2 & \ 5 & 2.68 & \ 3.84 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.04 \cr \ 2 & \ 5 & 2.68 & \ 3.84 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.36 \cr \ 2 & \ 5 & 2.68 & \ 3.84 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.36 \cr \ 2 & \ 5 & 2.68 & \ 3.84 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.70 \cr \noalign{\smallskip} \ 2 & \ 15 & 2.20 & \ 3.00 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.44 \cr \ 2 & \ 15 & 2.20 & \ 3.00 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.09 \cr \ 2 & \ 15 & 2.20 & \ 3.00 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.13 \cr \ 2 & \ 15 & 2.20 & \ 3.00 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.66 \cr \ 2 & \ 15 & 2.20 & \ 3.00 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 2.57 \cr \noalign{\smallskip} \ 2 & \ 30 & 1.93 & \ 2.57 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.87 \cr \ 2 & \ 30 & 1.93 & \ 2.57 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.22 \cr \ 2 & \ 30 & 1.93 & \ 2.57 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.85 \cr \ 2 & \ 30 & 1.93 & \ 2.57 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.87 \cr \ 2 & \ 30 & 1.93 & \ 2.57 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 2.77 \cr \noalign{\smallskip} \ 3 & \ 15 & 2.16 & \ 3.08 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.47 \cr \ 3 & \ 15 & 2.16 & \ 3.08 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.09 \cr \ 3 & \ 15 & 2.16 & \ 3.08 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.01 \cr \ 3 & \ 15 & 2.16 & \ 3.08 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.49 \cr \ 3 & \ 15 & 2.16 & \ 3.08 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 2.01 \cr \noalign{\smallskip} \ 3 & \ 30 & 1.90 & \ 2.63 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 2.07 \cr \ 3 & \ 30 & 1.90 & \ 2.63 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.30 \cr \ 3 & \ 30 & 1.90 & \ 2.63 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.63 \cr \ 3 & \ 30 & 1.90 & \ 2.63 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.72 \cr \ 3 & \ 30 & 1.90 & \ 2.63 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 2.12 \cr \noalign{\smallskip} \ 4 & \ 15 & 2.12 & \ 3.16 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.42 \cr \ 4 & \ 15 & 2.12 & \ 3.16 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.06 \cr \ 4 & \ 15 & 2.12 & \ 3.16 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.78 \cr \ 4 & \ 15 & 2.12 & \ 3.16 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.34 \cr \ 4 & \ 15 & 2.12 & \ 3.16 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.64 \cr \noalign{\smallskip} \ 4 & \ 30 & 1.88 & \ 2.68 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 2.02 \cr \ 4 & \ 30 & 1.88 & \ 2.68 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.26 \cr \ 4 & \ 30 & 1.88 & \ 2.68 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 2.26 \cr \ 4 & \ 30 & 1.88 & \ 2.68 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.55 \cr \ 4 & \ 30 & 1.88 & \ 2.68 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.71 \cr \noalign{\smallskip} \ 5 & \ 15 & 2.08 & \ 3.24 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.35 \cr \ 5 & \ 15 & 2.08 & \ 3.24 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.02 \cr \ 5 & \ 15 & 2.08 & \ 3.24 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.57 \cr \ 5 & \ 15 & 2.08 & \ 3.24 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.24 \cr \ 5 & \ 15 & 2.08 & \ 3.24 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.42 \cr \noalign{\smallskip} \ 5 & \ 30 & 1.85 & \ 2.74 & \ 5 & \hfil 0.4 & \hfil 0.6 & \hfil 1.88 \cr \ 5 & \ 30 & 1.85 & \ 2.74 & \ 10 & \hfil 0.0 & \hfil 0.5 & \hfil 1.19 \cr \ 5 & \ 30 & 1.85 & \ 2.74 & \ 10 & \hfil 0.4 & \hfil 0.6 & \hfil 1.94 \cr \ 5 & \ 30 & 1.85 & \ 2.74 & \ 10 & \hfil 0.2 & \hfil 0.8 & \hfil 1.41 \cr \ 5 & \ 30 & 1.85 & \ 2.74 & \ 20 & \hfil 0.4 & \hfil 0.6 & \hfil 1.47 \cr \noalign{\smallskip} \noalign{\hrule} }}$$ \end{table} For certain combinations of the characterizing parameters of the supernova remnant model, $E_{\rm ej}$, $M_{\rm ej}$ and $n_{\rm H0}$, the delayed decay of ionized $^{44}$Ti\ can have a sizable effect on the $\gamma$-ray activity of Cas~A which is observable at the present age of about 320 years. In Tables~1--3 the ratio ${\cal A}$ denotes the currently observable activity according to our model, normalized to the expected activity if ionization effects are not taken into account. The results depend on the assumed density enhancement $\alpha_{\rm clmp}$ in the clumps relative to the density of the homogeneous ejecta, and on the location of the clumps within a shell in the expanding ejecta bounded by the lower and upper mass coordinates $q_0$ and $q_1$, respectively. In addition, in Tables~1--3 numbers are given for the blast-wave radius $R_{\rm b}$ and the blast-wave velocity $v_{\rm b}$ as predicted by the McKee \& Truelove (1995) model at the current age of the Cas~A supernova remnant. The ratio ${\cal A}$ exhibits the following tendencies. For fixed explosion energy and ejecta mass, ${\cal A}$ increases with the density of the circumstellar gas of the supernova remnant because the expansion and dilution of the ejecta are slowed down for higher $n_{\rm H0}$ (i.e., $v_{\rm b}$ and $R_{\rm b}$ are smaller at the same time). An increase of the ratio ${\cal A}$ can also be seen when the explosion energy becomes larger but all other parameters are kept constant. The opposite trend is visible if the explosion energy and ambient density are fixed, in which case the ratio ${\cal A}$ decreases with larger ejecta mass. Although the blast-wave velocity at the present time is higher, it was lower early after the explosion and therefore the blast-wave radius is smaller for the larger ejecta mass. It is not easy to give simple explanations for these trends. The ratio ${\cal A}$, which scales with the current abundance of $^{44}$Ti\ and its effective decay rate, reflects the whole time evolution of the remnant and contains contributions from all parts of the shocked ejecta with their different ionization histories. The nonlinear dependence of ${\cal A}$ on the remnant parameters is a result of the interplay between a number of effects. For example, the post-shock density and temperature in the ejecta, in particular in the clumps, are important for the degree of ionization. The time when the reverse shock hits the clumps determines the time left for the equilibration between electrons and ions and for the duration of the delay of the $^{44}$Ti\ decay. This delay lasts from the moment of nearly complete ionization until recombination takes place again. In order to obtain a large effect on the $^{44}$Ti\ radioactivity that is measured in Cas~A presently, a high post-shock density and temperature are favorable for efficient ionization of $^{44}$Ti. On the other hand, the present-day $\gamma$-ray emission of Cas~A means that the temperature must not have decreased too slowly and the density not too rapidly. If that happened, $^{44}$Ti\ would be hindered from recombination and the observed $\gamma$-ray emission could not be explained. The maximum effect from the delayed decay of ionized $^{44}$Ti\ is found if most of the $^{44}$Ti-carrying clumps have been mixed roughly half-way into the ejecta, i.e., if they are assumed to be located in the $q$-interval between 0.4 and 0.6 (compare Fig.~2), and if the density enhancement in the clumps is around 10 for the higher explosion energies of $E_{\rm ej} = 2\times 10^{51}\,{\rm erg}$ and $3\times 10^{51}\,{\rm erg}$ (Tables~2 and 3, respectively). In case of the lowest considered value of the explosion energy, $E_{\rm ej} = 1\times 10^{51}\,{\rm erg}$ (Table~1), a clump over-density of about 5 seems preferable because of the higher density of the more slowly expanding ejecta. Higher than optimum $\alpha_{\rm clmp}$-values do not ensure rapid and strong ionization, whereas smaller than optimum $\alpha_{\rm clmp}$-values do not allow for quick recombination so that the decay rate stays low. If $^{44}$Ti\ is located very far out in the remnant (large $q$), the temperature behind the reverse shock is too low for efficient ionization (see Fig.~1). If $^{44}$Ti\ sits deep inside the ejecta (small $q$), the clumps have been reached by the reverse shock not sufficiently long ago so as to show a significant effect from $^{44}$Ti\ ionization. We found maximum values of ${\cal A}$ between roughly 1.5 and 2.5 for a large number of different combinations of remnant parameters and assumed clump locations and density enhancement factors. This means that the present-day $^{44}$Ti\ radioactivity of Cas~A could be 50\% up to more than a factor of 2 higher for a certain amount of $^{44}$Ti, if the decay half-life of $^{44}$Ti\ was stretched by ionization effects. Or, reversely, the current $\gamma$-ray emission of Cas~A due to $^{44}$Ti\ decay might be explained with a significantly lower production of this nucleus during the supernova explosion. Figure~4 shows that the relative effect from the retardation of the $^{44}$Ti\ decay will increase with the age of the remnant. For the different considered combinations of parameter values, Cas~A has reached slightly different stages of the remnant evolution around the transition time from the ejecta-dominated to the Sedov-Taylor phase. The tabulated blast-wave radius and velocity can be compared with observations which give values for the circumstellar density of about 20 hydrogen atoms per cm$^3$, an estimated ejecta mass of about $(2 \sim 5)\,$M$_{\odot}$ and a blast-wave radius of 2--3~pc at a distance of 3.4~kpc (see references given in Sect.~2). \section{Summary} We examined the effects of the reverse shock on the $^{44}$Ti\ $\beta$-decay rate in young supernova remnants. For this purpose, we employed the analytic remnant model of McKee \& Truelove (1995), which was used to describe the hydrodynamic evolution. We assumed that $^{44}$Ti\ is carried by $^{56}$Fe-dominated, dense clumps which are mixed into the otherwise homogeneous supernova ejecta and account for a minor fraction of the ejecta mass. Strong ionization of $^{44}$Ti\ due to the heating by the reverse shock and a sizable influence on the $^{44}$Ti\ decay is obtained if the clumps are assumed to be located at intermediate values of the radial mass coordinate $q$. The observationally important change of the $^{44}$Ti\ decay activity, however, depends also on the values of the remnant parameters, namely the explosion energy, the ejecta mass, the density of the ambient medium and the assumed overdensity factor of the clumps relative to the embedding homogeneous ejecta. \begin{figure} \epsfxsize=0.99\hsize \epsffile{ti44fig42.ps} \epsfxsize=0.99\hsize \epsffile{ti44fig41.ps} \caption{The decrease with time of the amount of $^{44}$Ti\ in the supernova remnant, $N(t)$, by $\beta$-decays, starting from the initial amount $N(0)$ ({\it right scale} in logarithm), and the corresponding, mostly increasing activity ratio ${\cal A}$ ({\it left scale}). Values of 3 M$_\odot$ ({\it upper panel}) and 5 M$_\odot$ ({\it lower}) are used for $M_{\rm ej}$, while $E_{\rm ej} = 2 \times 10^{51}$ erg, $n_{\rm H0} = 15$ cm$^{-3}$ and $\alpha_{\rm clmp} = 10$ are adopted in both cases. The different lines show results obtained for different assumptions of the radial locations of the clumps: $q = [q_0, q_1]$ = [0.0, 0.5] ({\it solid line}), [0.4, 0.6] ({\it dash-dotted line}), and [0.2, 0.8] ({\it dashed line}). The exponential decrease of the $^{44}$Ti\ abundance with the laboratory rate is plotted by the {\it dotted line} } \end{figure} We applied our model to the case of Cas~A, a young supernova remnant with an age of about 320 years. Observational data are considered to constrain the parameter space which was explored. We found that under certain conditions the ionization of $^{44}$Ti\ and the corresponding delay of its decay can yield an up to three times higher $^{44}$Ti\ activity at the present time than predicted on the grounds of the laboratory decay rate. This effect is large enough to reduce the apparent discrepancies between the $^{44}$Ti\ production in the explosion as inferred from the COMPTEL $\gamma$-line measurements and the theoretical expectations from the current supernova nucleosynthesis models. We emphasize again that to get an enhanced $^{44}$Ti\ decay activity we had to assume $^{44}$Ti\ to be associated with inhomogeneities in the ejecta which contain the Fe-group elements. Hydrodynamical models have so far not been able to predict the fraction of Fe and $^{44}$Ti\ in such clumpy structures, the overdensity of the clumps relative to the surrounding material at the considered age of a supernova remnant, the dynamics of the clumps, and their size and distribution. Multi-dimensional supernova models are called for, which connect the very early phase of the explosion with the remnant evolution a few hundred years later. Self-consistent multi-dimensional simulations have yet to be carried out for an improved theoretical picture of the explosive $^{44}$Ti\ production in supernovae. All published theoretical yields have so far only been obtained with ad-hoc assumptions, most often by spherically symmetric models and/or simulations of the supernova explosions which were started from artificial initial conditions (``piston'' models) rather than self-consistent situations. We add that such calculations also suffer from uncertainties concerning the nuclear reaction rates which are important to determine the $^{44}$Ti\ yield (The et al. 1998). On the other hand, more detailed X-ray and $\gamma$-ray observations with a better spatial (and energy) resolution are desirable to confirm or reject the implications of the described model, e.g., the existence of Fe-containing knots with temperatures in the approximate range of $(1 \sim 5)$ keV (cf. Tsunemi 1997). The current angular resolution of the ASCA/SIS observations, about one arcmin, is insufficient to measure the thermodynamic properties of individual clumps in the ejecta of Cas~A. Only information is available for the average electron temperature in the matter that is heated by the forward and reverse shocks. Comparing the observations with predictions from the remnant model used in this paper would therefore require monitoring of the thermodynamic history of the homogeneous ejecta. For this, one would need to compute the thermal state of the electrons, and would have to specify the unknown composition of the ejecta, which is very uncertain because it depends on the type of progenitor and explosion. The described investigations of $^{44}$Ti decay in supernova remnants could be done without such additional assumptions, but the models do not provide a description of the thermodynamic state of the remnant, which allows for a direct comparison with current observational data. Finally, we mention that we applied our analysis also to the new supernova remnant, RX~J0852.0-4622, discovered in the direction of the Vela remnant by Aschenbach (1998), in which a source of $^{44}$Ti\ line emission has been detected (Iyudin et al. 1998). However, for the estimated low density of the ambient medium of less than $0.04 (d / 500 {\rm pc})^{-0.5}$ cm$^{-3}$ with a distance $d$ to the remnant (Aschenbach 1998), we found that the reverse shock does not heat the ejecta to sufficiently high temperatures to ionize $^{44}$Ti. {\acknowledgements{ We thank H.~Tsunemi, J.~Vink, E.~Miyata and K.~Nomoto for helpful information and discussions. This work was supported by the SFB-375 ``Astroparticle Physics'' of the Deutsche Forschungsgemeinschaft. }} \vskip12pt \noindent {\bf Appendix A: reaction rates and related uncertainties} \vskip8pt \noindent We estimate here the $\beta$-decay, ionization and recombination rates, which appear in Eq.~(7). We also evaluate their uncertainties in relation to the results presented in Sect.~4. \vskip10pt {\it The orbital-electron capture rate of $^{44}${\rm Ti}}: We first formulate the $\beta$-decay rate of (near) neutral $^{44}$Ti. The corresponding half-life $t_{1/2}$ has been well determined to be 60 y (Ahmad et al. 1998; G\"orres et al. 1998; Norman et al. 1998; Wietfeldt et al. 1999). The $^{44}$Ti\ $\beta$-decay is by orbital-electron capture almost exclusively to the 146 keV excited ($0^-$) state of $^{44}$Sc (Endt 1998), the corresponding $Q$-value being 266 keV. We thus write the laboratory rate, $\lambda_{\rm lab} = {\rm ln}~2/t_{1/2}$, as \begin{equation} \lambda_{\rm lab} = \lambda_{\rm K} + \lambda_{\rm LI} + .... \end{equation} \noindent where $\lambda_{\rm K}, \lambda_{\rm LI}$ etc. refer to the rates for partial captures of K (1s$_\frac{1}{2}$), LI (2s$_\frac{1}{2}$) electrons etc. In the ``normal'' approximation (e.g. Konopinski \& Rose 1965), $\lambda_{\rm K}$ for example reads \begin{equation} \lambda_{\rm K} = \frac{G_{\rm A}^2}{4\pi^2} | M_0 |^2 (Q + B_{\rm K} )^2 g_{\rm K}^2 \, , \end{equation} \noindent for the transition of our current interest, where $G_{\rm A}$ is the axial-vector coupling constant of weak interaction, $M_0$ is a linear combination of two major nuclear matrix elements in the rank-0 first-forbidden ($0^+ \rightarrow 0^-$) transition, $Q$ is the $Q$-value, and $B_{\rm K}$ is the binding energy of the K-shell. Thus, $q_\nu \equiv Q + B_{\rm K}$ is the energy available to the emitted neutrino. (The nuclear recoil energy is very small and is ignored here.) Finally, $g_{\rm K}$ is the larger component of the radial wave-functions of the K-electron, which is evaluated at the nuclear radius. Since $Q$ is much larger than the binding energies (about 5 keV for K-electrons), the relative ratios of the partial decay rates become nearly equal to those of the respective wave-functions squared. Considering in addition that $\alpha Z $ is not very large, we have \begin{equation} \lambda_{\rm LI} \approx \frac{1}{8} \lambda_{\rm K}\, , \end{equation} since $g_{\rm LI}^2 \approx g_{\rm K}^2/8$ in the low-$\alpha Z$ approximation. With the use of the same approximation, one finds that the capture rates of the higher-shell electrons are small. Correspondingly, the orbital-electron capture rates of $^{44}$Ti\ with $i$ orbital electrons [i.e., (22-$i$)-times ionized] are approximately given by \begin{equation} \lambda_{\beta,i} \approx p_{i,{\rm K}} \lambda_{i,{\rm K}} + p_{i,{\rm LI}} \lambda_{i,{\rm LI}}\, , \end{equation} \noindent where $p_{i,{\rm K}}$ and $p_{i,{\rm LI}}$ are the occupancies ([0,1]) of the K and L$_{\rm I}$ orbits, respectively, and $\lambda_{i,{\rm K}}$ and $\lambda_{i,{\rm LI}}$ are the corresponding capture rates for the {\it filled} shells. We approximate the partial capture rates by \begin{equation} \lambda_{i,{\rm K}} \approx \frac{8}{9} \lambda_{i,{\rm K+LI}} \approx \frac{8}{9} \lambda_{\rm lab},\ \lambda_{i,{\rm LI}} \approx \frac{1}{9} \lambda_{i,{\rm K+LI}} \approx \frac{1}{9} \lambda_{\rm lab}\, , \end{equation} \noindent so that the total (effective) decay rate reads \begin{equation} \lambda_{\rm eff} \approx \Bigl( \frac{4}{9} n_1 + \frac{8}{9} n_2 + \frac{17}{18} n_3 + \sum_{i \geq 4} n_i \Bigr) \lambda_{\rm lab} \, {\bigg /}\, \sum_{i \geq 0} n_i \, , \end{equation} \noindent where $n_i$ is the number abundance of $^{44}$Ti$^{+22-i}$. The above procedure is quite well suited for dealing with the decays of highly-ionized $^{44}$Ti. In fact, our calculations based on a relativistic mean field method (Libermann et al. 1971) show that the overall errors associated with those approximations, including the neglect of the decrease of screening and the slightly altered energetics along with ionization, are small enough (less than 10 \%) to be ignored in the present work. \vskip10pt {\it The ionization rates of $^{44}${\rm Ti} and $^{56}${\rm Fe}}: The rates of electron-induced ionization are given by \begin{equation} \lambda_{{\rm ion},i} = n_{\rm e} \sum_{i \geq 1} \Bigl( p_{i,{\rm K}} \left\langle \sigma_{{\rm ion},i}^{\rm (K)} v \right\rangle + p_{i,{\rm LI}} \left\langle \sigma_{{\rm ion},i}^{\rm (LI)} v \right\rangle +... ... \Bigr)\, , \end{equation} \noindent where $n_{\rm e}$ is the electron number density and $\left\langle \sigma_{{\rm ion},i} v \right\rangle $ is the Maxwellian-average of the cross section (times the electron velocity $v$) for ionization of filled K-, LI-... shell electrons. As in Eq.~(16), $p_{i,{\rm K}}$ and $p_{i,{\rm LI}}$ stand for occupancies of the K- and L$_{\rm I}$ shells in the $i$-th ion. Our estimates of $\sigma_{{\rm ion},i}$ rely on the experimental data on K-shell ionization by electrons in the relevant energy range of $6 - 50$ keV (Long et al. 1990). We start with the classical Bethe-Mott-Massey formula (e.g., Powell 1976), which for the K-shell ionization from an ion with $i$ bound electrons reads \begin{equation} \sigma_{{\rm ion},i}^{\rm (K)} = [ \pi {\rm e}^4 / (E B_{i,\rm K}) ] Z_{\rm K} b_{i,{\rm K}} {\rm ln}[4 c_{i,{\rm K}} E / B_{i,{\rm K}}]\, , \end{equation} \noindent where $B_{i,{\rm K}}$ is the K-binding energy (taken positive), $E (\geq B_{i,{\rm K}})$ is the incident electron energy, and $Z_{\rm K} = 2$ is the number of K-electrons, and we treat $b_{i,{\rm K}}$ and $c_{i,{\rm K}}$ as adjustable parameters. It turns out that the existing experimental data for K-ionization of Ti by electrons in the relevant $E$ range of 6 - 50 keV (Long et al. 1990) can be well reproduced by the choice of \begin{equation} 1 / c_{22,{\rm K}} \approx a + ( 4 - a ) {\rm exp}[ d ( E / B_{22,{\rm K}} - 1 )]\, , \end{equation} with $ a = 0.55$, $d = 0.45$, and $b_{22,{\rm K}} = 0.445.$ For simplicity, we use the same parameter values for the higher shells as well as for highly-ionized cases. Aside from the trivial replacements of $Z_{\rm K}$ by $Z_{\rm LI}$ etc., we use the binding energies computed for each ion by the relativistic formalism, because the rates are sensitive to those thresholds. For $^{44}$Ti, their values in keV are, e.g., $B_{1,{\rm K}} = 6.6, B_{2,{\rm K}} = 6.2, B_{3,{\rm K}} = 6.1, B_{3,{\rm LI}} = 1.4, B_{4,{\rm K}} = 5.9, B_{4,{\rm LI}} = 1.3,$ in contrast to $B_{22,{\rm K}} = 5.0, B_{22,{\rm LI}} = 0.56.$ The same parameter values also reproduce the K-shell ionization cross-section data for Ni (Long et al. 1990) within the experimental uncertainties, while slightly (up to 20 \%) over-estimating the limited data points for Mn. When considered for Fe, this order of the uncertainty of the ionization rates influences the radioactivity under consideration in this work only at a negligible level. \vskip10pt {\it The recombination rates of $^{44}${\rm Ti} and $^{56}${\rm Fe}}: The recombination rates are given by \begin{equation} \lambda_{{\rm rec},i} = n_{\rm e} \sum_{i \geq 0} \Bigl( q_{i,{\rm K}} \left\langle \sigma_{{\rm rec},i}^{\rm (K)} v \right\rangle + q_{i,{\rm LI}} \left\langle \sigma_{{\rm rec},i}^{\rm (LI)} v \right\rangle + ... \Bigr)\, , \end{equation} \noindent where $q_{i} \equiv 1 - p_{i}$ are the vacancies of the respective shell prior to the recombination. We adopt the well-known formulae (Stobbe 1930) for the recombination cross-sections in the non-relativistic approximation for a pure Coulomb field. They are expressed in analytic forms in terms of the ratios of the incident electron energies ($E \geq 0$) to the binding energies of respective shells. The relativistic and screening effects can roughly be estimated with the use of effective charges, which turn out to be largely insignificant for the purpose of the present work.
\section{Introduction} To each real semisimple Jordan algebra, the Tits-Koecher-Kantor theory associates a distinguished parabolic subgroup $P=LN$ of a semisimple Lie group $G$. The groups $P$ which arise in this manner are precisely those for which $N$ is abelian, and $P$ is conjugate to its opposite $\overline{P}.$ Each non-open $L$-orbit ${\cal O}$ on $N^{*}$ admits an $L$-equivariant measure $d\mu $ which is unique up to scalar multiple. By Mackey theory, we obtain a natural irreducible unitary representation $\pi _{{\cal O}}$ of $P$% , acting on the Hilbert space \[ {\cal H}_{{\cal O}}=L^2({\cal O},d\mu ). \] In this context, we wish to consider two problems: \begin{enumerate} \item Extend $\pi _{{\cal O}}$ to a unitary representation of $G.$ \item Decompose the tensor products $\pi _{{\cal O}}\otimes \pi _{{\cal O}% ^{\prime }}\otimes \pi _{{\cal O}^{\prime \prime }}\otimes \cdots $ \end{enumerate} If the Jordan algebra is Euclidean (i.e. formally real) then $G/P$ is the Shilov boundary of a symmetric tube domain. In this case, the first problem was solved in \cite{sahi-expl}, \cite{shilov}, where it was shown that $\pi _{{\cal O}}$ extends to a unitary representation of a suitable covering group of $G$. The second problem was solved in \cite{tens}, where we established a correspondence between the unitary representations of $G$ occurring in the tensor product, and those of a ``dual'' group $G^{\prime }$ acting on a certain reductive homogeneous space. This correspondence agrees with the $\theta $-correspondence in various classical cases, and also gives a duality between $E_7$ and real forms of the Cayley projective plane. In this paper we start to consider these two problems for {\em non-Euclidean} Jordan algebras. The algebraic groundwork has already been accomplished in \cite{sahi-dp}, however the analytical considerations are much more subtle, and here we only treat the case of the representation $\pi _1=\pi _{{\cal O}% _1}$ corresponding to the {\em minimal} $L$-orbit ${\cal O}_1$. It turns out that in order for the first problem to have a positive solution, one has to exclude certain Jordan algebras of rank $2.$ This is related to the Howe-Vogan result on the non-existence of minimal representations for certain orthogonal groups. To each of the remaining Jordan algebras we attach a restricted root system $% \Sigma $ of rank $n$, where $n$ is the rank of the Jordan algebra. The root multiplicities, $d$ and $e$, of $\Sigma $ play a decisive role in our considerations. For the reader's convenience, we include a list of the corresponding groups $G$ and the multiplicities in the appendix. For these groups, we show that $\pi _1$ extends to a spherical unitary representation of $G$, and that the spherical vector is closely related to the {\em one}{\it \/} variable Bessel $K$-function $K_\tau (z)$, where \[ \tau =\frac{d-e-1}2. \] The function $K_\tau (z)$ can be characterized, up to a multiple, as the unique solution of the modified Bessel equation \[ \psi ^{\prime \prime }+z^{-1}\psi ^{\prime }-(1+\frac{\tau ^2}{z^2})\psi =0 \] that decays (exponentially) as $z\rightarrow \infty $; and, to us, one of the most delightful aspects of the present consideration is the unexpected and uniform manner in which this classical differential equation emerges from the structure theory of $G$. More precisely, we establish the following result: We identify $N$ with its Lie algebra ${\frak n}=\limfunc{Lie}(N)$ via the exponential map. We also fix an invariant bilinear form on $\left\langle \cdot ,\cdot \right\rangle $ on ${\frak g}$, which is a certain multiple of the Killing form, normalized as in Definition \ref{=form} below. We use this form to identify $N^{*}$ with $\overline{{\frak n}}=\limfunc{Lie}(\overline{N% })$. For $y$ in $\overline{{\frak n}}$, $\left\langle -\theta y,y\right\rangle $ is positive, and we define \[ |y|=\sqrt{\left\langle -\theta y,y\right\rangle }. \] \begin{theorem} $\pi _1$ extends to a unitary representation of $G$ with spherical vector $% |y|^{-\tau }K_\tau (|y|)$.\label{=Main-theorem} \end{theorem} Since $\pi _1$ is spherical, its Langlands parameter is its infinitesimal character, and this can be determined via the (degenerate) principal series imbedding described in section 2 below. It is then straightforward to verify that $\pi _1$ is the minimal representation of $G$, with annihilator equal to the Joseph ideal. (For $G=GL(n),$ the minimal representation is not unique.) Thus our construction should be compared to other realizations of the minimal representations in \cite{brylinski}, \cite{torasso}, \cite{huang} etc. Although our construction is for a more restrictive class of groups, it does offer two advantages over the other constructions. The first advantage is that our construction works for a larger class of representations, and the second advantage is that it is well-suited for tensor product computations. Both of these features will be explored in detail in a subsequent paper. In the present paper, we consider $k$-fold tensor powers of $\pi _1$, where $k$ is strictly smaller than $n$ (rank of $\Sigma $), and show that the decomposition can be understood in terms of certain reductive homogeneous spaces \[ G_k/H_{k\text{ }},1<k<n. \] These spaces are defined in section 3, and are listed in the appendix. We consider also the corresponding Plancherel decomposition: \[ L^2(G_k/H_k)=\int_{\widehat{G}_k}^{\oplus }m(\kappa )\kappa \,d\mu (\kappa )\,\text{,} \] where $d\mu $ is the Plancherel measure, and $m(\kappa )$ is the multiplicity function. Then we have \begin{theorem} For $1<k<n$, there is a correspondence $\theta _k$ between $\widehat{G}_k$ and $\widehat{G},$ such that \[ \pi _1^{\otimes k}=\int_{\widehat{G}_k}^{\oplus }m(\kappa )\theta _k\left( \kappa \right) d\mu (\kappa ). \] \label{=Main-tensoring} \end{theorem} \section{Preliminaries} The results of this section are all well-known. Details and proofs may be found in \cite{sahi-expl}, \cite{kostant-sahi} and in the references therein (in particular, \cite{braun} and \cite{loos}). \subsection{Root multiplicities} Let $G$ be a real simple Lie group and let $K$ be a maximal compact subgroup corresponding to a Cartan involution $\theta .$ We shall denote the Lie algebras of $G$, $K$ etc by ${\frak g}$, ${\frak k}$ etc. Their complexifications will be denoted by lowercase fraktur letters with subscript $_{{\Bbb C}}$. Fix $\theta $, and let ${\frak g}={\frak k}+{\frak p% }$ be the associated Cartan decomposition. The parabolic subgroups $P=LN$ obtained by the Tits-Kantor-Koecher construction are those such that $N$ is abelian, and $P$ is $G$-conjugate to its opposite parabolic \[ \overline{P}=\theta (P)=L\overline{N}. \] In this case $N$ has a natural structure of a real Jordan algebra, which is unique up to a choice of the identity element. In (Lie-)algebraic terms, this means that $P$ is a maximal parabolic subgroup corresponding to a simple\ (restricted) root $\alpha $ which has coefficient $1$ in the highest root, and which is mapped to $-\alpha $ under the long element of the Weyl group. In this situation, $M:=K\cap L$ is a symmetric subgroup of $K$ (this is {\em % equivalent} to the abelianness of $N),$ and we fix a maximal toral subalgebra ${\frak t}$ in the orthogonal complement of ${\frak m}$ in $% {\frak k}$. The roots of ${\frak t}_{{\Bbb C}}$ in ${\frak g}_{{\Bbb C}}$ form a restricted root system of type $C_n$, where $n=\dim _{{\Bbb R}}{\frak t}$ is the (real) rank of $N$ as a Jordan algebra (this result is essentially due to C. Moore). We fix a basis $\{\gamma _1,\gamma _2,\ldots ,\gamma _n\}$ of $% {\frak t}^{*}$ such that \[ \Sigma ({\frak t}_{{\Bbb C}},{\frak g}_{{\Bbb C}})=\{\pm (\gamma _i\pm \gamma _j)/2,\pm \gamma _j\}\text{.} \] The restricted root system $\Sigma =\Sigma ({\frak t}_{{\Bbb C}},{\frak k}_{% {\Bbb C}})$ is of type $A_{n-1},C_n$ or $D_n$, and the first of these cases arises precisely when $N$ is a Euclidean Jordan algebra. This case was studied in \cite{sahi-expl}, therefore we restrict our attention to the last two cases. The root multiplicities in $\Sigma $ play a key role in our considerations. If $\Sigma $ is $C_n$, there are two multiplicities, corresponding to the short and long roots, which we denote by $d$ and $e,$ respectively. If $% \Sigma $ is $D_n$, and $n\neq 2,$ then there is a single multiplicity, which we denote by $d$, so that $D_n$ may be regarded as a special case of $C_n$, with $e=0.$ The root system $D_2$ is reducible (being isomorphic to $A_1\times A_1$) and {\em a priori} there are two root multiplicities. In what follows, we explicitly{\em \ }exclude the case when these multiplicities are different.% {\em \ }This means that we exclude from consideration the groups \[ G=O(p,q),N={\Bbb R}^{p-1,q-1}(p\neq q); \] indeed, our main results are false for these groups.{\em \ }When the two multiplicities {\em coincide}, we once again denote the common multiplicity by $d$. The multiplicity of the short roots $\pm (\gamma _i\pm \gamma _j)/2$ in $% \sum ({\frak t}_{{\Bbb C}},{\frak g}_{{\Bbb C}})$ is equal to $2d$, and the multiplicity of the long roots $\pm \gamma _i$ is $e+1$. In the appendix we include a table listing the groups under consideration, as well as the values of $d$ and $e$ for each of these groups. \subsection{Cayley transform} We briefly review the notion of the Cayley transform. Let $C$ be the following element (of order $8$) in $SL_2\left( {\Bbb C}\right) $ \[ C=\frac 1{\sqrt{2}}\left( \begin{array}{cc} 1 & i \\ i & 1 \end{array} \right) . \] The Cayley transform of ${\frak sl}_2({\Bbb C})$ is the automorphism (of order $4$) given by \[ c=\text{Ad }C. \] It transforms the ``usual'' basis of ${\frak sl}_2({\Bbb C})$ \[ x=\left( \begin{array}{cc} 0 & 1 \\ 0 & 0 \end{array} \right) ,\text{ }y=\ \left( \begin{array}{cc} 0 & 0 \\ 1 & 0 \end{array} \right) ,\text{ }h=\left( \begin{array}{cc} 1 & 0 \\ 0 & -1 \end{array} \right) , \] to the basis \[ X=\frac 12\left( \begin{array}{cc} -i & 1 \\ 1 & i \end{array} \right) ,\text{ }Y=\frac 12\left( \begin{array}{cc} i & 1 \\ 1 & -i \end{array} \right) ,\text{ }H=i\left( \begin{array}{cc} 0 & -1 \\ 1 & 0 \end{array} \right) , \] where $X=c(x)=C^{-1}xC,$ etc. In turn, $c$ can be expressed as \[ c=\exp \text{ ad }\frac{\pi i}4(x+y)=\exp \text{ ad }\frac{\pi i}4(X+Y). \] The key property of the Cayley transform is that it takes the compact torus (spanned by $iH)$ to the split torus spanned by $h$ (cf. \cite{koranyi-wolf}% ). We turn now to the Lie algebra ${\frak g}_{{\Bbb C}}$. By the Cartan-Helgason theorem the root spaces ${\frak p}_{\gamma _j}$ are one-dimensional, and so by the Jacobson-Morozov theorem we get holomorphic homomorphisms \[ \Phi _j:{\frak sl}_2({\Bbb C})\longrightarrow {\frak g}_{{\Bbb C}}\text{, }% j=1,...,n \] such that $X_j=\Phi _j(X)$ spans ${\frak p}_{\gamma _j}$. We fix such maps $\Phi _j$, and denote the images of $x,X,y,Y,h,H$ by $% x_j,X_j,$ etc. Since the roots $\gamma _j$ are strongly orthogonal, the triples $\{X_j,Y_j,H_j\}$ commute with each other, and the Cayley transform of ${\frak g}$ is defined to be the automorphism \[ c=\exp \text{ ad }\frac{\pi i}4\left( \sum X_j+\sum Y_j\right) =\prod \exp \text{ ad }\frac{\pi i}4(X_j+Y_j). \] Thus we obtain an ${\Bbb R}$-split toral subalgebra ${\frak a}$ defined by \[ {\frak a}=c^{-1}(i{\frak t})={\Bbb R}h_1\oplus \cdots \oplus {\Bbb R}h_n. \] The roots of ${\frak a}_{{\Bbb C}}$ in ${\frak g}_{{\Bbb C}}$ are \[ \Sigma ({\frak a}_{{\Bbb C}},{\frak g}_{{\Bbb C}})=\left\{ \pm \varepsilon _i\pm \varepsilon _j,\pm 2\varepsilon _j\right\} \text{ where }\varepsilon _i=\frac 12\gamma _i\ \circ c. \] The short roots have multiplicity $2d$ and the long roots have multiplicity $% e+1.$ In fact ${\frak a}\subset {\frak l},$ and we have \[ \Sigma ({\frak a},{\frak l})=\left\{ \pm (\varepsilon _i-\varepsilon _j)\right\} ,\text{ }\Sigma ({\frak a},{\frak n})=\left\{ \varepsilon _i+\varepsilon _j,2\varepsilon _j\right\} ,\text{ }\Sigma ({\frak a},% \overline{{\frak n}})=\left\{ -\varepsilon _i-\varepsilon _j,-2\varepsilon _j\right\} \] \begin{definition} The invariant form $\left\langle .,.\right\rangle $ on ${\frak g}$ is normalized by requiring \[ \left\langle x_1,y_1\right\rangle =1. \] \label{=form} \end{definition} For $y\in \overline{{\frak n}}$, we set $\left| y\right| \stackrel{\text{def}% }{=}\sqrt{-\left\langle y,\theta y\right\rangle }$ , as in Introduction. \subsection{ Orbits and measures} We now describe the orbits of $L$ in $\overline{{\frak n}}$ $\simeq N^{*}.$ For $k=1,...,n-1,$ define \[ {\cal O}_k=L\cdot (y_1+y_2+\ldots +y_k). \] Then these, together with the trivial orbit ${\cal O}_0$, comprise the totality of the singular (i.e., non-open) $L$-orbits in $\overline{{\frak n}} $. We define $\nu \in {\frak a}^{*}$ as \[ \nu =\varepsilon _1+\varepsilon _2+\ldots +\varepsilon _n.\text{ } \] Then $\nu $ extends to a character of ${\frak l}$ , and we will write $e^\nu $ for the corresponding (spherical) character of $L.$ \begin{lemma} The orbit ${\cal O}_1$ carries a natural $L-$equivariant measure $d\mu _1$, which transforms by the character $e^{2d\nu }$, that is \[ \int_{{\cal O}_1}g(l\cdot y)d\mu _1(y)=e^{2d\nu }(l)\int_{{\cal O}% _1}g(y)d\mu _1(y)\text{.} \] \end{lemma} \TeXButton{Proof}{\proof} Let $S_1$ be the stabilizer of $y_1$ in $L$. It suffices to show that the modular function of$\,$ $S_1$ is the restriction, from $L$ to $S_1$, of the character $e^{2d\nu }$. Passing to the Lie algebra ${\frak s}_1$, we need to show that \[ \limfunc{tr}\limfunc{ad}\nolimits_{{\frak s}_1}=2d\nu |_{{\frak s}_1}. \] To see this, we remark that ${\frak s}_1$ has codimension $1$ inside a maximal parabolic subalgebra ${\frak q}$ of ${\frak l}$, corresponding to the stabilizer of the line through $y_1.$ The space of characters of ${\frak % q}$ is two-dimensional, and it follows that the space of characters of $% {\frak s}_1$ is one-dimensional. Hence any character of ${\frak s}_1$ is determined by its restriction to ${\frak a}\cap {\frak s}_1=\limfunc{Ker}% \varepsilon _1$. The restriction of $\nu $ to ${\frak s}_1$ is nontrivial, hence \[ \limfunc{tr}\limfunc{ad}\nolimits_{{\frak s}_1}=k\nu \] for some constant $k$. Obviously, $\limfunc{tr}\limfunc{ad}_{{\frak l}}=0$, and the only root spaces missing from ${\frak s}_1$ are the root spaces ${\frak l}% _{\varepsilon _1-\varepsilon _j}$, $j\geq 2$ (each of these root spaces has dimension $2d$). Hence, for $a\in {\frak a}$% \[ \limfunc{tr}\limfunc{ad}\nolimits_{{\frak s}_1}(a)=-2d\sum_{j=2}^n(% \varepsilon _1-\varepsilon _j)(a)\text{,} \] and restricting this to $\limfunc{Ker}\varepsilon _1$, we obtain $2d\nu \,|_{% {\frak a}\cap {\frak s}_1}$.\TeXButton{End Proof}{\endproof} \smallskip\ {\em Example.} Consider $G=O_{2n,2n}$ realized as the group of all $2n\times 2n$ real matrices preserving the split symmetric form $\left( \begin{array}{cc} 0 & I_{2n} \\ I_{2n} & 0 \end{array} \right) $. Then $P=LN=GL_{2n}({\Bbb R})\rightthreetimes \limfunc{Skew}_{2n}(% {\Bbb R}).$ More precisely, \[ L=\left\{ \left( \begin{array}{cc} A & 0 \\ 0 & A^{t^{-1}} \end{array} \right) :A\in GL_{2n}({\Bbb R})\right\} \] and \[ N=\left\{ \left( \begin{array}{cc} I_{2n} & 0 \\ B & I_{2n} \end{array} \right) :B+B^t=0\right\} . \] Then \[ {\frak a}=\left\{ \limfunc{diag}(a_1,a_1,a_2,a_2,\ldots ,a_n,a_n,-a_1,-a_1,-a_2,-a_2,\ldots ,-a_n,-a_n),\,a_i\in {\Bbb R}\right\} \] is the toral subalgebra of ${\frak g}$ (and ${\frak l}$) described in the preceding subsection. We can take \[ y_1=\left( \begin{array}{cc} 0_{2n} & B_1 \\ 0 & 0_{2n} \end{array} \right) \text{, where }B_1=\left( \begin{array}{ccc} 0 & -1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 0 \end{array} \right) . \] The Lie algebra ${\frak s}_1$ of the stabilizer $S_1=\limfunc{Stab}_Ly_1$ can be written as \[ {\frak s}_1=\left\{ \left( \begin{array}{cc} A & 0 \\ 0 & -A^t \end{array} \right) :A=\left( \begin{array}{cc} A_{11} & 0 \\ A_{21} & A_{22} \end{array} \right) ,A_{11}\in {\frak sl}_2,A_{22}\in {\frak gl}_{2n-2}\right\} \] It is a codimension 1 subalgebra of the parabolic subalgebra ${\frak q}$ of $% {\frak gl}_{2n}$, where ${\frak q}=({\frak gl}_2+{\frak gl}_{2n-2})+{\Bbb R}% ^{2,2n-2}.$\medskip\ {\em Remark.} In this example $\nu =\frac 12\limfunc{tr}$, $d=2$ and $% e^{2d\nu }=(\det )^2$. \section{Minimal representation of $G$} If $\chi $ is a character of ${\frak l}$, we write $\pi _\chi $ for the (unnormalized) induced representation $\limfunc{Ind}_{\overline{P}}^G(\chi )$% . These representations were studied in \cite{sahi-dp} in the ``compact'' picture, by algebraic methods. Among the results established there was the existence of a finite number of ``small'', unitarizable, spherical subrepresentations, which occur for the following values of $\chi $ \[ \chi _j=e^{-jd\nu },\text{ }j=1,\ldots ,n-1. \] In this paper we use analytical methods, and work primarily with the ``non-compact'' picture, which is the realization of $\pi _\chi $ on $% C^\infty (N)$, via the Gelfand-Naimark decomposition \[ G\approx N\overline{P}. \] In fact, using the exponential map we can identify ${\frak n}$ and $N$, and realize $\pi _\chi $ on $C^\infty ({\frak n}).$ We will show that the unitarizable subrepresentation of $\pi _{\chi _1}$ admits a natural realization on the Hilbert space $L^2({\cal O}_1,d\mu )$. Since there is no obvious action of $G$ on this space, we have to proceed in an indirect fashion. The key is an explicit realization of the spherical vector $\sigma _{\chi _1}.$ \subsection{The Bessel function} We let $d,e$ be the root multiplicities of $\Sigma ({\frak t},{\frak k})$ as in previous section, and define \[ \tau _G=\tau =(d-e-1)/2 \] as in the introduction. Let $K_\tau $ be the $K$-Bessel function on $(0,\infty )$ satisfying \begin{equation} z^2K_\tau ^{\prime \prime }+zK_\tau ^{\prime }-(z^2+\tau ^2)K_\tau =0. \label{=bessel-eq} \end{equation} Put $\phi _\tau (z)=\dfrac{K_\tau (\sqrt{z})}{\left( \sqrt{z}\right) ^\tau }$% , then $\phi _\tau $ satisfies the differential equation \begin{equation} D\phi _\tau =0,\text{where }D\phi =4z\phi ^{\prime \prime }+4(\tau +1)\phi ^{\prime }-\phi \text{.} \label{=change} \end{equation} We lift $\phi _\tau $ to an $M$-invariant function $g_\tau $ on ${\cal O}_1$% , by defining \begin{equation} g_\tau (y)=\phi _\tau (-\left\langle y,\theta y\right\rangle )=\frac{K_\tau \left( \left| y\right| \right) }{\left| y\right| ^\tau }. \label{=g-tau} \end{equation} {\em Remark}. If $d=e$ (as is the case for $G=Sp_{2n}({\Bbb C})$ or $% Sp_{n,n} $), then $\tau =-\frac 12$ and \[ g_\tau (y)=\left| y\right| ^{1/2}K_{-1/2}(\left| y\right| )=\left| y\right| ^{1/2}\frac{\exp (-\left| y\right| )}{\left| y\right| ^{1/2}}=e^{-\left| y\right| }. \] If $d=e+1$ (this is true for $GL_{2n}({\bf k})$, ${\bf k}={\Bbb R}$, ${\Bbb C% }$ or ${\Bbb H}$), then \[ g_\tau (y)=K_0(\left| y\right| ). \] \begin{proposition} {\em (1)} $g_\tau $ is a (square-integrable) function in $L^2({\cal O}% _1,d\mu _1)$. {\em (2) }The measure $g_\tau d\mu _1$ defines a tempered distribution on $% \overline{{\frak n}}$.\label{=sq-int} \end{proposition} \TeXButton{Proof}{\proof} (1) We define \[ {\cal O}^{\prime }\stackrel{\text{def}}{=}\{y^{\prime }\in {\cal O}_1:\left| y^{\prime }\right| =1\}. \] Then ${\cal O}^{\prime }$ is compact; the map \[ {\cal O}^{\prime }\times (0,\infty )\ni (y^{\prime },w)\longmapsto wy^{\prime }\in {\cal O}_1 \] is a diffeomorphism, and the measure $d\mu _1$ can be decomposed as a product \[ d\mu _1(wy^{\prime })=d\mu ^{\prime }(y^{\prime })d\mu ^{\prime \prime }(w) \] We now determine the explicit form of $d\mu ^{\prime \prime }(w).$ Define $h=\sum_{i=1}^nh_i$, then $(\limfunc{ad}h)y=-2y$ for any $y\in \overline{{\frak n}}$. We take $y\in {\cal O}_1$, $z>0$, $a=\ln z$ and calculate \[ d\mu _1(zy)=d\mu _1(\exp (-a\frac h2)\cdot y)=e^{-2d\nu (-\frac{ah}2)}d\mu _1(y)=e^{dna}d\mu _1(y)=z^{dn}d\mu _1(y)\text{.} \] Therefore, for $z>0$ \begin{equation} d\mu _1(zy)=z^{dn}d\mu _1(y) \label{=scaling} \end{equation} and it follows that $d\mu ^{\prime \prime }(zw)=z^{dn}d\mu ^{\prime \prime }(w),$ and so, up to a scalar multiple, \[ d\mu ^{\prime \prime }(w)=w^{dn-1}dw, \] where $dw$ is the Lebesgue measure. We can now calculate \begin{eqnarray} \int_{{\cal O}_1}\left| g_\tau (y)\right| ^2d\mu _1(y) &=&\int_0^\infty \int_{{\cal O}^{\prime }}\frac{K_\tau (w)^2}{w^{2\tau }}d\mu ^{\prime }(y^{\prime })w^{dn-1}dw \nonumber \\ &=&c\int_0^\infty \frac{K_\tau (w)^2}{w^{2\tau }}w^{dn-1}dw, \label{=sep-var} \end{eqnarray} where $c=\mu ^{\prime }({\cal O}^{\prime })$ is a positive constant. The function $K_\tau (w)$ has a pole of order $\tau $ at $0$ (or, in case of $% \tau =0,$ a logarithmic singularity at $0$), and it decays exponentially as $% w\rightarrow \infty $ \cite[3.71.15]{watson}. Hence $w^{-2\tau }K_\tau (w)^2$ has a pole of order \[ 4\tau =2(d-e-1)\leq 2d-2<dn-1 \] (recall that we require $n\geq 2)$. Thus the integrand in (\ref{=sep-var}) is non-singular and decays exponentially as $w\rightarrow \infty $. Therefore, the integral (\ref{=sep-var}) converges and $g_\tau (y)\in L^2(% {\cal O}_1,d\mu _1)$. (2) From the calculation in (1), we see that $g_\tau (y)\in L_{\text{loc}}^1(% {\cal O}_1,d\mu _1)$ and has exponential decay at $\infty $ (i.e., as $% \left| y\right| \rightarrow \infty $). This implies the result. \TeXButton{End Proof}{\endproof}\smallskip We can now define the Fourier transform of $g_\tau $, \[ \Phi =\widehat{g_\tau d\mu _1} \] as a (tempered) distribution on ${\frak n.}$ The key result is the following \begin{proposition} $\Phi $ is a multiple of the spherical vector $\sigma _{\chi _1}$.\label {=spher} \end{proposition} The proof of this proposition will be given over the next two subsections.% \label{=analytic} \subsection{Characterization of spherical vectors} For $\phi :{\frak n}\rightarrow {\frak n},$ let $\xi (\phi )$ denote the corresponding vector field: \[ \xi (\phi )f(x)=\left. \frac d{dt}f(x+t\phi (x))\right| _{t=0}\text{ for }f:% {\frak n}\rightarrow {\Bbb C}. \] Then we have the following formulas for the action of $\pi _\chi $ on $% C^\infty ({\frak n})$: \begin{itemize} \item for $x_0\in {\frak n}$, $\pi _\chi (x_0)=\xi (x_0)$, \item for $h_0\in {\frak l}$, $\pi _\chi (h_0)=\chi (h_0)-\xi \left( [h_0,x]\right) $, \item for $y_0\in \overline{{\frak n}}$, $\pi _\chi (y_0)=\chi [x,y_0]-\frac 12\xi \left( [h,x]\right) $, where $h=[x,y_0]$. \end{itemize} We need a Lie algebra characterization of $\sigma _\chi $: \begin{lemma} The space of $\pi _\chi ({\frak k})$-invariant distributions on ${\frak n}$ is 1--dimensional (and spanned by $\sigma _\chi $). \end{lemma} \TeXButton{Proof}{\proof} It is well known (and easy to prove) that the only distributions on ${\Bbb R}^n,$ which are annihilated by $\frac \partial {\partial x_i},$ $i=1,...,n$ are the constants. More generally, we can replace ${\Bbb R}^n$ by a manifold, and $\left\{ \frac \partial {\partial x_i}\right\} $ by any set of vector fields which span the tangent space at each point of the manifold. For $\chi =0$, the formulas above show that $\pi _0({\frak g})$ acts by vector fields on $C^\infty ({\frak n}).$ Moreover, using the decomposition $% G=K\overline{P},$ we see that $\pi _0({\frak k})$ is a spanning family of vector fields. Thus the result follows in this case. For general $\chi $, if $T$ is a $\pi _\chi ({\frak k})$-invariant distribution, then $T/\sigma _\chi =T\sigma _{-\chi }$ is $\pi _0({\frak k})$% -invariant, and hence a constant. \TeXButton{End Proof}{\endproof} \begin{proposition} Let $T$ be an $M$-invariant distribution on ${\frak n}$ such that \[ \pi _\chi (y+\theta y)T=0\,\text{ for {\em some} }y\text{ }\neq 0\text{ in }% \overline{{\frak n}}, \] then $T$ is a multiple of the spherical vector $\sigma _\chi $.\label{=infin} \end{proposition} \TeXButton{Proof}{\proof} The $M$-invariance of $T$ implies that \[ \pi _\chi ({\frak m})T=0 \] Since ${\frak m}$ is a maximal subalgebra of ${\frak k}$, ${\frak m}$ and $% y+\theta y$ generate ${\frak k}$ as a Lie algebra. Thus \[ \pi _\chi ({\frak k})T=0, \] and the result follows from the previous lemma. \TeXButton{End Proof} {\endproof} \subsection{The $K$-invariance of the Bessel function} We now turn to the proof of Proposition \ref{=spher}. To simplify notation, we will write $\pi $ instead of $\pi _{\chi _1}.$ Since $\Phi $ is clearly $% M $-invariant, by Proposition \ref{=infin} it suffices to show \[ \pi (y_1+\theta y_1)\Phi =0 \] for $y_1\in \overline{{\frak n}}$. We will prove this through a sequence of lemmas. It is convenient to introduce the following notation: if $g_1$ and $g_2$ are functions on ${\cal O}_1$, we define \[ (g_1,g_2)=\int_{{\cal O}_1}g_1(y)\,g_2(y)d\mu _1(y) \] provided the integral converges. If $g$ is a function on ${\cal O}_1$ and $h\in {\frak l}$, then the action of $h$ on $g$ is given by \[ h\cdot g(y)\stackrel{\text{def}}{=}\left. \frac d{dt}g(e^{th}\cdot y)\right| _{t=0}. \] In the computation below, we shall work with the expressions of the type \[ \left. \left( \frac d{dt}\int_{{\cal O}_1}g(e^{th}\cdot y)d\mu (y)\right) \right| _{t=0}. \] To justify differentiation under the integral sign, we need to impose the standard conditions on $g$ (e.g. \cite[p.170]{kestleman}), as follows. Define a class of functions ${\cal I}\subset C^\infty ({\cal O}_1)$, given by the following conditions: a smooth function $g\ $belongs to ${\cal I}$ if \begin{itemize} \item $g\in L^1({\cal O}_1,d\mu _1)$ and \item for any $h\in {\frak l}$ we can find $c>0$ and $G(y)\in L^1({\cal O}% _1,d\mu _1)$, such that \[ \left| \left. \frac d{dt}g(e^{th}\cdot y)\right| _{t=t_0}\right| \leq G(y) \] for all $y\in {\cal O}_1$ and $\left| t_0\right| <c$. \end{itemize} \begin{lemma} Suppose $g_1,g_2$ are smooth functions on ${\cal O}_1$, such that $g_1g_2\in {\cal I}$. Then \begin{equation} (h\cdot g_1,g_2)+(g_1,h\cdot g_2)=2d\nu (h)(g_1,g_2). \label{=leib} \end{equation} \label{=leibnitz} \end{lemma} \TeXButton{Proof}{\proof} Using the $L$-equivariance of $d\mu _1$, we obtain \[ \int_{{\cal O}_1}g_1(e^{th}y)g_2(e^{th}y)\,d\mu _1=e^{2td\nu (h)}\int_{{\cal % O}_1}g_1g_2\,d\mu _1. \] Under the assumptions of the lemma, we can differentiate this identity in $t$% , to get \[ \int_{{\cal O}_1}h\cdot (g_1g_2)\,d\mu _1=2d\nu (h)\int_{{\cal O}% _1}g_1g_2\,d\mu _1\text{.} \] By the Leibnitz rule, the result follows.\TeXButton{End Proof}{\endproof} More generally, if $g_1,g_2$ are functions on ${\frak n}\times {\cal O}_1$, then $(g_1,g_2)$ is a function on ${\frak n}$. In this notation, for $g$ in $% L^1({\cal O}_1,d\mu _1)$, the Fourier transform of $gd\mu _1$ is given by the formula \[ \widehat{gd\mu _1}=(e^{-i\left\langle x,y\right\rangle },g). \] \begin{lemma} Let $g\in L^1({\cal O}_1,d\mu _1)$ be a smooth function on ${\cal O}_1$, such that \[ e^{-i\left\langle x,y\right\rangle }g\in {\cal I}. \] Suppose $f=(e^{-i\left\langle x,y\right\rangle },g)$, then \[ \pi (y_1)f=-\frac 12(e^{-i\left\langle x,y\right\rangle },h\cdot g(y))\text{% , where }h=[x,y_1]\text{.\label{=pi-y0}} \] \end{lemma} \TeXButton{Proof}{\proof} By the formula for the action of $\pi (y_1),$ we get \begin{eqnarray*} -2\left( \pi (y_1)f+d\nu (h)f\right) &=&\xi \left( [h,x]\right) \cdot (e^{-i\left\langle x,y\right\rangle },g) \\ &=&\frac d{dt}\left. \left( e^{-i\left\langle x+t[h,x],y\right\rangle },g\right) \right| _{t=0} \\ &=&\frac d{dt}\left. \left( e^{-i\left\langle x,y-t[h,y]\right\rangle },g\right) \right| _{t=0} \\ &=&-\left( h\cdot e^{-i\left\langle x,y\right\rangle },g\right) \\ &=&(e^{-i\left\langle x,y\right\rangle },h\cdot g)-2d\nu (h)(e^{-i\left\langle x,y\right\rangle },g). \end{eqnarray*} where we have used the previous lemma, and the relation \[ \left\langle x+t[h,x],y\right\rangle =\left\langle x,y\right\rangle +t\left\langle [h,x],y\right\rangle =\left\langle x,y\right\rangle -t\left\langle x,[h,y]\right\rangle =\left\langle x,y-t[h,y]\right\rangle . \] The result follows.\TeXButton{End Proof}{\endproof} The pairing $-\left\langle \cdot ,\theta \cdot \right\rangle $ gives a positive definite $M$-invariant inner product on $\overline{{\frak n}}$, and we now obtain the following \begin{lemma} Suppose that $g(y)=\phi \left( -\left\langle y,\theta y\right\rangle \right) $ for some smooth $\phi $ on $\left( 0,\infty \right) $, and $% e^{-i\left\langle x,y\right\rangle }g\in {\cal I}$. Put $f(x)=(e^{-i\langle x,y\rangle },g)$, as before. Then \[ \pi (y_1)f=\left( e^{-i\left\langle x,y\right\rangle },\left\langle x,[[\theta y,y_1],y]\right\rangle \phi ^{\prime }\left( -\left\langle y,\theta y\right\rangle \right) \right) . \] \end{lemma} \TeXButton{Proof}{\proof} Writing $h=[x,y_1]$ as in the previous lemma, we get \begin{eqnarray*} h\cdot g(y) &=&\frac d{dt}\left. \phi \left( -\left\langle y+t[h,y],\theta (y+t[h,y])\right\rangle \right) \right| _{t=0} \\ &=&\frac d{dt}\left. \phi \left( -\left\langle y,\theta y\right\rangle -2t\left\langle \theta y,[h,y]\right\rangle +O(t^2)\right) \right| _{t=0} \\ &=&-2\left\langle \theta y,[h,y]\right\rangle \phi ^{\prime }\left( -\left\langle y,\theta y\right\rangle \right) \text{.} \end{eqnarray*} Since \[ \left\langle \theta y,[h,y]\right\rangle =\left\langle \theta y,[[x,y_1],y]\right\rangle =\left\langle x,[[\theta y,y_1],y]\right\rangle , \] the result follows.\TeXButton{End Proof}{\endproof} The key lemma is the following computation \begin{lemma} Let $\phi $ and $f$ be as in the previous lemma, and suppose for $x\in {\frak n}$ \begin{equation} e^{-i\left\langle x,y\right\rangle }\phi \left( -\left\langle y,\theta y\right\rangle \right) \in {\cal I}\text{, \ \ }e^{-i\left\langle x,y\right\rangle }\phi ^{\prime }\left( -\left\langle y,\theta y\right\rangle \right) \in {\cal I}\text{.} \label{=diff-cond} \end{equation} Then we have \begin{equation} \pi (y_1+\theta y_1)f(x)=\left( e^{-i\left\langle x,y\right\rangle },i\left\langle \theta y_1,y\right\rangle \,(D\phi )\left( -\left\langle y,\theta y\right\rangle \right) \right) , \label{=crown} \end{equation} where the differential operator $D$ is given by the formula {\em (\ref {=change})}, i.e. \begin{equation} (D\phi )\left( -\left\langle y,\theta y\right\rangle \right) =4\left( -\left\langle y,\theta y\right\rangle \right) \phi ^{\prime \prime }+2(d+1-e)\phi ^{\prime }-\phi \text{.} \label{=de1} \end{equation} \label{=main-mess} \end{lemma} \TeXButton{Proof}{\proof} Choose a basis $l_j$ of ${\frak l}$ and define functions $c_j(y)$ by the formula $[\theta y,y_1]=\sum_jc_j(y)l_j$. Then by the previous lemma \begin{eqnarray*} \pi (y_1)f &=&\sum_j(e^{-i\left\langle x,y\right\rangle },\left\langle x,[l_j,y]\right\rangle c_j\phi ^{\prime })=i\sum_j\frac d{dt}\left. \left( e^{-i\left\langle x,y+t[l_j,y]\right\rangle },c_j\phi ^{\prime }\right) \right| _{t=0} \\ &=&i\sum_j(l_j\cdot e^{-i\left\langle x,y\right\rangle },c_j\phi ^{\prime })% \text{.} \end{eqnarray*} Differentiation in this calculation is justified, because $e^{-i\left\langle x,y\right\rangle }\phi ^{\prime }\left( -\left\langle y,\theta y\right\rangle \right) \in {\cal I}$. Applying (\ref{=leib}) to the last expression, we can write \begin{equation} \pi (y_1)f=-i\sum_j\left( e^{-i\left\langle x,y\right\rangle },-2d\nu (l_j)c_j\phi ^{\prime }+c_jl_j\cdot \phi ^{\prime }+\phi ^{\prime }l_j\cdot c_j\right) \text{.} \label{=pi-y1} \end{equation} We now calculate each term in this expression. \begin{itemize} \item First we have \[ \sum_j\nu (l_j)c_j\phi ^{\prime }=\nu ([\theta y,y_1])\phi ^{\prime }. \] Since $\nu $ is a real character of ${\frak l}$, it vanishes on ${\frak l}% \cap {\frak k}$ and we have $\nu ([\theta y,y_1])=\nu ([\theta y_1,y])$. Recall that ${\frak n}$ and $\overline{{\frak n}}$ are irreducible ${\frak l} $-modules. Therefore, $\nu ([\theta y_1,y])=k\left\langle \theta y_1,y\right\rangle $ for some constant $k\neq 0,$ independent of $y$. Setting $y=y_1$, we get $\left\langle \theta y_1,y_1\right\rangle =\left\langle -x_1,y_1\right\rangle =-1$. Hence $k=-\nu ([\theta y_1,y_1])=1$% , and therefore \begin{equation} -\sum_j2d\nu (l_j)c_j\phi ^{\prime }=-2d\left\langle \theta y_1,y\right\rangle \phi ^{\prime }\text{.} \label{=term1} \end{equation} \item Next we compute \begin{eqnarray*} \sum_jc_jl_j\cdot \phi ^{\prime }=[\theta y,y_1]\cdot \phi ^{\prime } \\ =\frac d{dt}\left. \phi ^{\prime }\left( -\left\langle y+t[[\theta y,y_1],y],\theta (y+t[[\theta y,y_1],y])\right\rangle \right) \right| _{t=0} \\ =-2\left\langle y,[[y,\theta y_1],\theta y]\right\rangle \phi ^{\prime \prime }\left( -\left\langle y,\theta y\right\rangle \right) \end{eqnarray*} Since $y$ is a ${\frak k}$-conjugate to a root vector, there is a scalar $% k^{\prime }$ independent of $y$ such that $[[y,\theta y],y]=k^{\prime }\left\langle y,\theta y\right\rangle y$ . Setting $y=y_1$ we get $% \left\langle y_1,\theta y_1\right\rangle =-1$, \[ [[y_1,-x_1],y_1]=-2y_1 \] and $k^{\prime }=2.$ Also $-\left\langle y,[[y,\theta y_1],\theta y]\right\rangle =\left\langle [y,\theta y],[y,\theta y_1]\right\rangle =\left\langle [[y,\theta y],y],\theta y_1\right\rangle $. Hence \begin{equation} \sum_jc_jl_j\cdot \phi ^{\prime }=4\left\langle y,\theta y\right\rangle \left\langle \theta y_1,y\right\rangle \phi ^{\prime \prime }\text{.} \label{=term2} \end{equation} \item Next we note that $\sum_jl_j\cdot c_j$ is independent of the basis $% l_j$, so we may assume that \[ \theta l_j=\pm l_j\text{ and }\left\langle l_j,-\theta l_k\right\rangle =\delta _{jk}\text{.} \] Then $c_j(y)=\left\langle [\theta y,y_1],-\theta l_j\right\rangle $ and \begin{eqnarray*} \tsum\nolimits_jl_j\cdot c_j=\tsum\nolimits_j\left\langle [\theta [l_j,y],y_1],-\theta l_j\right\rangle \\ =\tsum\nolimits_j\left\langle y_1,[\theta [l_j,y],\theta l_j]\right\rangle =-\left\langle y_1,\Omega \theta y\right\rangle . \end{eqnarray*} Here $\Omega =\tsum\nolimits_j\limfunc{ad}(\theta l_j)^2=\Omega _{{\frak l}% }-2\Omega _{{\frak k}\cap {\frak l}}\,$, where the Casimir elements are obtained by using dual bases with respect to $\left\langle \,,\,\right\rangle $. To continue, we need the following lemma: \end{itemize} \begin{lemma} $\Omega $ acts on ${\frak n}$ by the scalar $k^{\prime \prime }=2-2e$.\label {kpp} \end{lemma} \TeXButton{Proof}{\proof} When $e=1$ it's easy to see that the operator $% \Omega $ acts by $0$. Indeed, \ in this case ${\frak g}$ is a complex semisimple Lie algebra and for each basis element $l_j\in {\frak k}\cap {\frak l}$ there exists a basis element $l_j^{\prime }=\sqrt{-1}l_j\in {\frak p}\cap {\frak l}$ . Then $[l_j,[l_j,x]]+[l_j^{\prime },[l_j^{\prime },x]]=0$ and $k^{\prime \prime }=0$. When $e=0,$ ${\frak g}$ is split and simply laced, and ${\frak l}$ is the split real form of a complex reductive algebra ${\frak l}_{{\Bbb C}}$. Take a root vector $x_\lambda \in {\frak g}_\lambda $, where $\lambda $ is any positive root in ${\frak n}$. For any positive root $\alpha $ of ${\frak l}_{% {\Bbb C}}$ we fix $e_\alpha \in {\frak l}_\alpha $ and set $l_\alpha =e_\alpha +\theta e_\alpha \in {\frak k}\cap {\frak l}$ and $l_\alpha ^{\prime }=e_\alpha -\theta e_\alpha \in {\frak p}\cap {\frak l}$. Then the collection of all $l_\alpha $, $l_\alpha ^{\prime }$ together with the orthonormal basis of a Cartan subalgebra ${\frak f}$ of ${\frak l}$ forms a basis of ${\frak l}$. Observe that \[ \lbrack l_\alpha ,[l_\alpha ,x_\lambda ]]+[l_\alpha ^{\prime },[l_\alpha ^{\prime },x_\lambda ]]=[e_\alpha ,[e_\alpha ,x_\lambda ]]+[e_{-\alpha },[e_{-\alpha },x_\lambda ]]=0\text{,} \] since $x_\lambda \in {\frak g}_\lambda $ and neither $\lambda +2\alpha $ nor $\lambda -2\alpha $ is a root of the simply laced algebra ${\frak g}_{{\Bbb C% }}$. We choose a basis $\{u_i\}$ of ${\frak f}$, and denote the elements of the dual (with respect to $\left\langle \;,\;\right\rangle $) basis by $% \widetilde{u}_i.$ Then \[ \Omega x_\lambda =\sum_i[u_i,[\widetilde{u}_i,x_\lambda ]]=\left\langle \lambda ,\lambda \right\rangle x_\lambda =2x_\lambda \text{.} \] In the remaining two cases ${\frak k}\cap {\frak l}$ acts on ${\frak n}$ irreducibly, therefore $\Omega $ automatically acts by a scalar and it suffices to compute $\sum_j[l_j,[l_j,x_1]]$. For $e=3$ we have $G=GL_{2n}(% {\Bbb H}),L=GL_n({\Bbb H})\times GL_n({\Bbb H})$ and ${\frak n}={\Bbb H}% ^{n\times n}$. The computation for this group is similar to the case of $% G=GL_{2n}({\Bbb R})$. We reduce the calculation to the summation over the diagonal subalgebra of ${\frak l}$ and obtain \[ \Omega x_\lambda =\left\langle \lambda ,\lambda \right\rangle x_\lambda +3\left\langle \sqrt{-1}\lambda ,\sqrt{-1}\lambda \right\rangle x_\lambda =-4x_\lambda \text{.} \] Finally, for $e=2$ $(G=Sp_{n,n})$, a direct evaluation of $% \sum_j[l_j,[l_j,x_1]]$ gives $k^{\prime \prime }=-2$.\TeXButton{End Proof} {\endproof} \smallskip\ \\\ Therefore, we get \begin{equation} \sum_j\phi ^{\prime }l_j\cdot c_j=-2(1-e)\left\langle \theta y_1,y\right\rangle \phi ^{\prime }\text{.} \label{=term3} \end{equation} \begin{itemize} \item Finally, we have \begin{equation} \pi (\theta y_1)f=\frac d{dt}\left. \left( e^{-i\left\langle x+t\theta y_1,y\right\rangle },\phi \right) \right| _{t=0}=-i\left( e^{-i\left\langle x,y\right\rangle },\left\langle \theta y_1,y\right\rangle \phi \right) . \label{=term4} \end{equation} \smallskip\ \end{itemize} Putting the formulas (\ref{=term1})--(\ref{=term4}) together, we deduce the lemma.\TeXButton{End Proof}{\endproof}\smallskip\ {\bf Proof of Proposition \ref{=spher}.} Recall that we study $\phi _\tau (z)=\dfrac{K_\tau (\sqrt{z})}{\left( \sqrt{z}\right) ^\tau }$ , its lift $% g_\tau $ to the radial function on ${\cal O}_1$, \[ g_\tau (y)=\phi _\tau (-\left\langle y,\theta y\right\rangle )=\dfrac{K_\tau (\left| y\right| )}{\left| y\right| ^\tau } \] and its Fourier transform $\Phi (x)=(e^{-i\left\langle x,y\right\rangle },g_\tau )$. By Proposition \ref{=infin} it suffices to check that $\pi (y_1+\theta y_1)\Phi =0$. This identity would follow immediately from Lemma \ref{=main-mess}, because $D\phi _\tau =0$ by formula (\ref{=change}) and then the desired result follows from (\ref{=crown}). To complete the proof we have to verify the assumptions (\ref{=diff-cond}). In subsection \ref{=analytic} we proved that $g_\tau \in L^1({\cal O}_1,d\mu _1)$. It is easy to verify (using the standard facts about the derivatives of $K_\tau $ from \cite{watson}), that the lifts to ${\cal O}_1$ of the functions $\phi _\tau ^{\prime }(z)$ and $\phi _\tau ^{\prime \prime }(z)$ (we denote them by $g_\tau ^{\prime }(y)$ and $g_\tau ^{\prime \prime }(y)$) both belong to $L^1({\cal O}_1,d\mu _1)$. Observe also that $\phi _\tau (z)$% , $\phi _\tau ^{\prime }(z)$, $\phi _\tau ^{\prime \prime }(z)$ are all monotone on $(0,\infty )$. Moreover, since all these functions tend to zero exponentially as $\left| y\right| \rightarrow \infty $, the functions $A(y)g_\tau (y)$, $A(y)g_\tau ^{\prime }(y)$, $A(y)g_\tau ^{\prime \prime }(y)$ all belong to $L^1({\cal O}% _1,d\mu _1)$, for any $A(y)$ bounded in the neighbourhood of $y=0$ and growing (at most) polynomially with respect to $\left| y\right| $ as $\left| y\right| \rightarrow \infty $. Fix $h\in {\frak l}$ , $x\in {\frak n}$ and choose $c>0$ sufficiently small, such that for all $y\in {\cal O}_1$ and $\left| t\right| <c$% \[ \left| \left\langle e^{th}\cdot y,\theta (e^{th}\cdot y)\right\rangle \right| \geq \frac{\left| \left\langle y,\theta y\right\rangle \right| }2% \text{.} \] We can then estimate the derivative: \[ \left| \frac d{dt}\left( e^{-i\left\langle x,e^{th}\cdot y\right\rangle }\phi _\tau \left( -\left\langle e^{th}y,\theta e^{th}y\right\rangle \right) \right) \right| \leq \left| A_1(y)\phi _\tau \left( \left| y\right| ^2/2\right) \right| +\left| A_2(y)\phi _\tau ^{\prime }\left( \left| y\right| ^2/2\right) \right| \text{,} \] for all $y\in {\cal O}_1$ and $\left| t\right| <c$, where $A_1(y),A_2(y)$ are some functions of polynomial growth. From the discussion above, the right-hand side of this inequality is an $L^1$-function on ${\cal O}_1$, hence $e^{-i\left\langle x,y\right\rangle }g_\tau \in {\cal I}$. Proceeding in the same manner, we deduce that $e^{-i\left\langle x,y\right\rangle }g_\tau ^{\prime }\in {\cal I}$. \TeXButton{End Proof} {\endproof} \smallskip\ \subsection{Proof of Theorem 0.1} Denote by ${\bf J}$ the space of the induced representation $\pi _1=\limfunc{% Ind}_{\overline{P}}^G(e^{-d\nu })$. By the Gelfand-Naimark decomposition and the $\exp $ map, ${\bf J}$ can be viewed as a subspace of $C^\infty ({\frak n% })$. Then for $l\in L$ and $\eta \in {\bf J}$ we have \[ \pi _1(l)\eta (x)=e^{-d\nu }(l)\eta (l^{-1}\cdot x)\text{.} \] It was proved in \cite{sahi-dp} that the $({\frak g},K)$-module ${\bf J}$ has a unitarizable spherical $({\frak g},K)$-submodule $V$, which we also regard as a subspace of $C^\infty ({\frak n})$. {\bf Remark. }It is possible to give a direct description of the elements of the ``abstract'' Hilbert space ${\cal H}$, where ${\cal H}$ is the Hilbert space closure of $V$ with respect to the $({\frak g},K)$-invariant norm on $% V $. For that purpose we use the ``compact'' realization of $\pi _1$ on $% C^\infty (K/M)$ from \cite{sahi-dp}. It was shown that $\pi _1$ is a representation of ladder type, with all its $K$-types $\{\alpha _m\,|\;m\in {\Bbb N}\}$ lying on a single line, $\alpha _1$ being a one-dimensional $K$% -type. The restriction $\left\langle \;,\;\right\rangle _m$ of a $\pi _1$% -invariant Hermitian form to any $K$-type $\alpha _m$ is a multiple of the $% L^2(K)$-inner product on $V$, and from the explicit formulas in \cite {sahi-dp} it follows that \[ q_m\stackrel{\text{def}}{=}\frac{\left\langle \;,\;\right\rangle _m}{% \left\langle \;,\;\right\rangle _1}=O(m^C) \] for some constant $C>1$, which can be expressed in terms of parameters $d,$ $% e$ and $n$. Thus we can identify ${\cal H}$ with the Hilbert space $% L^2\left( {\Bbb N},\{q_m\}\right) $, where the constant $q_m$ gives the weight of the point $m\in {\Bbb N}$. That is, any element of ${\cal H}$ can be viewed as an $M$-equivariant function on $K$, such that its sequence of Fourier coefficients belongs to $L^2\left( {\Bbb N},\{q_m\}\right) $. In particular $L^2\left( {\Bbb N},\{q_m\}\right) \subset l^2({\Bbb N)}$, and the elements of ${\cal H}$ all lie in $L^2(K)$.\smallskip\ We write ${\bf H}$ for the space of those tempered distributions on ${\frak n% }$ which are Fourier transforms of $\psi d\mu _1$ for some $\psi \in L^2(% {\cal O}_1,d\mu _1)$. If $\eta $ is the Fourier transform of a distribution of the form $\psi d\mu _1$, i.e., \[ \eta (x)=\int_{{\cal O}_1}e^{-i\left\langle x,y\right\rangle }\psi (y)d\mu _1(y)=\left( e^{-i\left\langle x,y\right\rangle },\psi (y)\right) , \] then \begin{eqnarray*} \pi _1(l)\eta (x) &=&e^{-d\nu }(l)\eta (l^{-1}\cdot x)=\int_{{\cal O}% _1}e^{-i\langle l^{-1}\cdot x,y\rangle }\psi (y)e^{-d\nu }(l)d\mu _1(y) \\ &=&\int_{{\cal O}_1}e^{-i\langle l^{-1}x,l^{-1}y\rangle }\psi (l^{-1}\cdot y)e^{-dv}(l)d\mu _1(l^{-1}\cdot y) \\ &=&(e^{-i\left\langle x,y\right\rangle },e^{d\nu }(l)\psi (l^{-1}\cdot y)). \end{eqnarray*} It follows from the calculation above that $P$ acts {\em unitarily }on{\em \ }${\bf H}$ (it is convenient to identify this action with its realization on $L^2({\cal O}_1,d\mu _1)$ via the Fourier transform).We denote this unitary representation of $P$ by $\pi ^{\prime }$. Observe that $(\pi ^{\prime },% {\bf H})$ is an irreducible representation of $P$. According to Proposition \ref{=sq-int}, $\Phi (x)=(e^{-i<x,y>},\left| y\right| ^{-\tau }K_\tau (\left| y\right| ))$ belongs to ${\bf H}$. \begin{theorem} $V$ is a dense subspace of ${\bf H}$, and the restriction of the norm is $(% {\frak g},K)$-invariant. \end{theorem} \TeXButton{Proof}{\proof} Let $C^\infty (K)_V$ be the subspace of $C^\infty (K)$, consisting of those smooth functions on $K$, whose $K$-isotypic components belong to $V$. Since $V$ is a submodule of ${\bf J}$, $C^\infty (K)_V$ is obviously $G$-invariant. Denote by ${\cal C}(G)$ the convolution algebra of smooth $L^1$ functions on $G=PK$, and consider \[ {\bf W}=\pi _1({\cal C}(G))\Phi \subset C^\infty (K)_V. \] So all elements of ${\bf W}$ are continuous functions on $K$, hence continuous on $G$, and therefore are determined by their restrictions to $N$% . Moreover, ${\bf W}=\pi _1({\cal C}(PK))\Phi $ and $K$ fixes $\Phi $, therefore \[ {\bf W}=\pi _1({\cal C}(P))\Phi =\pi ^{\prime }({\cal C}(P))\Phi \text{.} \] This shows that ${\bf W}$ is a $\pi ^{\prime }(P)$-invariant subspace of $% {\bf H}$, and from the irreducibility of $\pi ^{\prime }$ we conclude that $% {\bf W}$ is dense in ${\bf H}$. We can now put two $\pi _1(P)$-invariant norms on ${\bf W}$ -- one from $% {\bf H}$ and another from $V$, as follows. If $f=\sum c_mv_m$, with $v_m$ in the $K$-isotypic component with highest weight $\alpha _m$ (occurring in $V$% ) and $\left\| v_m\right\| _{L^2(K)}=1$, then \begin{equation} \left\| f\right\| _V^2=\sum \left| c_m\right| ^2q_m\text{.} \label{=new-norm} \end{equation} Since $f$ \thinspace is smooth, it follows that $\left| c_m\right| $ decays rapidly, so the series in (\ref{=new-norm}) converges, thus giving a $\pi _1(P)$-invariant norm on ${\bf W}$. Then it follows from \cite{poguntke} (cf. \cite[p.417]{sahi-expl}), that we can find a (dense) ${\cal C}(P)$-invariant subspace ${\bf W}^{\prime }\subset {\bf W}$, such that these two forms are proportional on ${\bf W}% ^{\prime }$. Considering the closure of ${\bf W}^{\prime }$ we obtain an isometric $P$-invariant imbedding of ${\bf H}$ into ${\cal H}$. Then ${\bf W}$ is: (1) \ a $G$-invariant subspace of the irreducible module ${\cal H}$, hence dense in ${\cal H}$; (2) \ a dense subspace of the Hilbert space ${\bf H}$. It follows that ${\bf H}={\cal H}$. \TeXButton{End Proof}{\endproof}% \smallskip\ This concludes the proof of Theorem \ref{=Main-theorem}. \section{Tensor powers of $\pi _1$} \subsection{Restrictions to $P$} In the previous section we constructed a unitary representation $\pi _1$ of $% G$ acting on the Hilbert space $L^2({\cal O}_1,d\mu _1)$, where ${\cal O}_1$ is the minimal $L$-orbit in a non-Euclidean Jordan algebra $N$. Define the $% k $-th tensor power representation \[ \Pi _k=\pi _1^{\otimes k}(2\leq k<n). \] As we shall show, the techniques developed in \cite{tens} allow us to establish a duality between the spectrum of this tensor power and the spectrum of a certain homogeneous space. We omit the proofs of the several propositions below, because the proofs of the corresponding statements from \cite{tens} can be used without any substantial modification. Observe that the orbit ${\cal O}_k$ is dense in $\stackunder{k\text{ times}}{% \underbrace{{\cal O}_1+{\cal O}_1+\ldots +{\cal O}_1}}$. The representation $% \Pi _k$ acts on $\left[ L^2({\cal O}_1,d\mu _1)\right] ^{\otimes k}\simeq L^2({\cal O}_k^{\prime },d\mu ^{\prime })$, where ${\cal O}_k^{\prime }=% {\cal O}_1^{\times k}$ and $d\mu ^{\prime }$ is the product measure on${\cal % \ O}_k^{\prime }$ . We fix a generic representative $\xi ^{\prime }=(\xi _1,\xi _2,\ldots ,\xi _k)\in {\cal O}_k^{\prime }$ , such that \[ \xi =\xi _1+\xi _2+\ldots +\xi _k\in {\cal O}_k. \] Denote by $S_k^{\prime }$ and $S_k$ the isotropy subgroups of $\xi ^{\prime } $ and $\xi $, respectively, with respect to the action of $L$ on ${\cal O}% _k^{\prime }$ and ${\cal O}_k$. Observe that the Lie algebras ${\frak s}% _k^{\prime }$ and ${\frak s}_k$ of $S_k^{\prime }$ and $S_k$, respectively, can be written as \begin{eqnarray*} {\frak s}_k^{\prime } &=&({\frak h}_k+{\frak l}_k)+{\frak u}_k \\ {\frak s}_k &=&({\frak g}_k+{\frak l}_k)+{\frak u}_k\text{.} \end{eqnarray*} Here ${\frak l}_k,{\frak g}_k$ and ${\frak h}_k$ are reductive, ${\frak h}% _k\subset {\frak g}_k$ and ${\frak u}_k$ is a nilpotent radical common for both ${\frak s}_k^{\prime }$ and ${\frak s}_k$. Let $G_k$ and $H_k$ be the corresponding Lie groups.\smallskip\ {\bf Example. }Take $G=O_{2n,2n}$ and $k<n$. Then $\xi _i=E_{2i-1,2i}-E_{2i,2i-1}$ ($1\leq i\leq k$), $\xi =\sum_{i=1}^k\xi _i$ and \[ {\frak s}_k=\left( {\frak sp}_{2k}({\Bbb R})+{\frak gl}_{2(n-k)}({\Bbb R)}% \right) +{\Bbb R}^{2k,2(n-k)}\text{.} \] Then $G_k=Sp_{2k}({\Bbb R)}$ and it's easy to check that $H_k=SL_2({\Bbb R}% )^k$.\smallskip\ The following Lemma can be verified by direct calculation (cf. \cite[Lemma 2.1]{tens}). \begin{lemma} Let $\chi _\xi $ be the character of $N$ corresponding to $\xi \in N^{*}$. Then \[ \Pi _k|_P=\limfunc{Ind}\nolimits_{S_k^{\prime }N}^P(1\otimes \chi _\xi )=% \limfunc{Ind}\nolimits_{S_kN}^P\left( (\limfunc{Ind}\nolimits_{S_k^{\prime }}^{S_k}1)\otimes \chi _\xi \right) \;\;\;\text{{\em (}}L^2\text{{\em % -induction).}\TeXButton{End Proof}{\endproof}} \] \end{lemma} Let $\gamma ^{\prime }=\limfunc{Ind}_{H_k}^{G_k}1$ be the quasiregular representation of $G_k$ on $L^2(G_k/H_k)$; then it can be decomposed using the Plancerel measure $d\mu $ for the reductive homogeneous space $% X_k=G_k/H_k$ and the corresponding multiplicity function $m:\widehat{G}% _k\rightarrow \{0,1,2,\ldots \}$, i.e., \[ \gamma ^{\prime }\simeq \int_{\widehat{G}_k}^{\oplus }m(\kappa )\kappa \,d\mu (\kappa )\text{.} \] Each irreducible representation $\kappa $ of $G_k$ can be extended to an irreducible representation $\kappa ^{\vee }$ of $S_k$, and the decomposition of the Lemma above can be rewritten as \begin{equation} \Pi _k|_P=\int_{\widehat{G}_k}^{\oplus }m(\kappa )\Theta (\kappa )\,d\mu (\kappa )\text{,} \label{=Pi-to-P} \end{equation} where $\Theta (\kappa )=\limfunc{Ind}_{S_kN}^P(\kappa ^{\vee }\otimes \chi _\xi )$. Moreover, by Mackey theory all representations $\Theta (\kappa )$ are unitary irreducible representations of $P$, and $\Theta (\kappa ^{\prime })\simeq \Theta (\kappa ^{^{\prime \prime }})$ if and only if $\kappa ^{\prime }\simeq \kappa ^{\prime \prime }$. \subsection{Low-rank theory} In \cite{tens} we extended the theory of low-rank representations (\cite {li-singular}) to the conformal groups of euclidean Jordan algebras. Inspection of the argument in \cite{tens} shows that the analogous theory can be developed in exactly the same manner for the conformal groups of {\em % non-euclidean} Jordan algebras. For any unitary representation $\eta $ of $G$, we decompose its restriction $% \eta |_N$ into a direct integral of unitary characters, where the decomposition is determined by a projection-valued measure on $\widehat{N}% =N^{*}$. If this measure is supported on a single {\em non-open} $L$% -orbit\thinspace {\em \ }${\cal O}_m$,\thinspace $1\leq m<n$ we call $\eta $ a {\em low-rank representation}, and write $\limfunc{rank}\eta =m$. Proceeding by induction on $m$, as in \cite{li-singular}, \cite[Sect 3]{tens}% , we can prove the following \begin{theorem} Let $\eta $ be a low-rank representation of $G.$ Write ${\cal A}(\eta ,P)$ for the von Neumann algebra generated by $\{\eta (x)|\,x\in P\}$ and ${\cal A% }(\eta ,G)$ for the von Neumann algebra generated by $\{\eta (x)|\,x\in G\}$% . Then ${\cal A}(\eta ,G)={\cal A}(\eta ,P)$. \TeXButton{End Proof} {\endproof}\smallskip\ \end{theorem} {\bf Proof of Theorem \ref{=Main-tensoring}. }Now consider the restriction of $\Pi _k$ to $N$. Its restriction to $P$ is given by the direct integral decomposition (\ref{=Pi-to-P}), and we can further restrict it to $N$. The rank of the induced representation $\Theta (\kappa )=\limfunc{Ind}% _{S_kN}^P(\kappa ^{\vee }\otimes \chi _\xi )$ is $k$ (the $N$--spectrum is supported on the $L-$orbit of $\xi $, i.e. ${\cal O}_k$). Therefore $\Pi _k$ can be decomposed over the irreducible representations of $G$ of rank $k$. It follows from the theorem above that any two non-isomorphic representations from the spectrum of $\Pi _k$ restrict to non-isomorphic irreducible representations of $P$. Hence the representation $\Pi _k$ can be decomposed as \begin{equation} \Pi _k=\int_{\widehat{G}_k}^{\oplus }m(\kappa )\theta (\kappa )\,d\mu (\kappa ), \label{=Pi-decomp} \end{equation} where for almost every $\kappa $ the unitary irreducible representation $% \theta (\kappa )$ is obtained as the {\em unique }irreducible representation of $G$ determined by the condition $\theta (\kappa )|_P=\Theta (\kappa )$. Therefore, the map $\kappa \rightarrow \theta (\kappa )$ gives a (measurable) bijection between the spectrum of $\Pi _k=\pi ^{\otimes k}$ and the unitary representations of $G_k$ occurring in the quasiregular representation on $L^2(G_k/H_k).$\TeXButton{End Proof}{\endproof}\smallskip\ {\bf Example. }Take $G=E_{7(7)}$. It is the conformal group of the split exceptional real Jordan algebra $N$ of dimension 27. Consider the tensor square of the minimal representation $\pi _1$ of $G$ ($k=2$). Then $L={\Bbb R% }^{*}\times E_{6(6)}$, $S_2^{\prime }$ is the stabilizer of $\,y_1$ {\bf and }$y_2$ and $S_2$ is the stabilizer of $y_1+y_2\in {\cal O}_2$. One can see that in this case ${\frak g}_2=\limfunc{Stab}_{{\frak s0}(5,5)}(y_1+y_2)=% {\frak so}(4,5)$ and ${\frak h}_2=\limfunc{Stab}_{{\frak s0}(5,5)}(y_1)\cap \limfunc{Stab}_{{\frak s0}(5,5)}(y_2)={\frak so}(4,4)$ (cf. \cite[16.7] {adams}). Hence the decomposition (\ref{=Pi-decomp}) establishes a duality between the representations of $E_{7(7)}$ occurring in $\Pi _2=\pi _1\otimes \pi _1$ and the unitary representations of $Spin(4,5)$ occurring in $% L^2\left( Spin(4,5)/Spin(4,4)\right) $ . The homogeneous space $% Spin(4,5)/Spin(4,4)$ is a (pseudo-riemannian) symmetric space of rank 1, and it is known to be multiplicity free. Therefore, $\pi _1\otimes \pi _1$ has simple spectrum. Similarly, for $G=E_7({\Bbb C})$ we obtain a duality between $E_7({\Bbb C})$ and the symmetric space $SO_9({\Bbb C})/SO_8({\Bbb C})$.\newpage\ \
\section{Introduction} Superstring theory has been regarded for a long time as the most promising quantum theory of gravity (see \cite{wit,pol}). Indeed it naturally includes the graviton among its massless states. Nevertheless it has intrinsic limits deriving from its own definition. As it is well known, this theory describes one dimensional extended objects (closed or open strings) vibrating in a ten dimensional Minkowsky space--time, whose spectrum of vibrational modes are expected to correspond to the observed fields in nature. From a mathematical point of view the theory is a $\sigma$--model defined on the $1+1$ world sheet of a string propagating in a $10$--dimensional target space. A dimensionful parameter characterizing the string and defining the scale of its vibrational levels is the string length $\ell_s$, whose square will be denoted by $\alpha^\prime$.\par Consistency of the theory requires its local invariance with respect to diffeomorphism and conformal transformations on the world sheet coordinates and metric. Among the massless fields in the spectrum of the theory there is a scalar $\phi$ (the dilaton) and a spin--$2$ field $G_{\mu \nu}$ which is identified with the graviton field. One of the vacuum solutions of the theory is the one on which the symmetric tensor gives the Minkowsky metric on the target space--time ($\langle G_{\mu \nu}\rangle=\eta_{\mu\nu}$) and $\langle \phi \rangle =\phi_0$. Superstring theory is quantized on this vacuum, perturbatively with respect to an effective coupling constant $g=Exp(\phi_0)$. Versions of the theory on certain other vacua of the form $M_d\times K$ representing compactifications of the ten dimensional target space to a lower dimensional Minkowsky space $M_d$ ($d\le 10$) times a suitable compact Ricci--flat manifold $K$ have been constructed as well. The low energy limit of these theories is described by a {\it $d$--dimensional supergravity}. On the other hand a second quantized formulation of superstring theory on a curved background characterized by a non Ricci--flat solution of the low--energy effective supergravity theory, is not known so far (except for very special cases) and represents in general a complicate problem (\cite{ktsetse}). As we shall see, one of the potential achievements of the duality recently conjectured by Maldacena (see \cite{malda},\cite{duff} and references therein), is to shed light on the quantum spectrum of superstring theory realized on a particular kind of non Ricci--flat vacuum, namely of the form $AdS\times K$, where $AdS$ stands for a $d$--dimensional Anti de Sitter space--time and $K$ is a suitable compact $(10-d)$--dimensional Einstein space. These vacua correspond to the geometry of some regions of space--time in presence of extended objects which are believed to belong to the non perturbative spectrum of superstring theories and therefore may be regarded as non--perturbative vacua of these theories. In the remaining part of the present introduction I shall try to further formalize these concepts in a pedagogical fashion, preparing a physical scenario in which to set Maldacena's conjecture.\par One of the great achievements of the last ten years is the concept of {\it duality}: correspondences have been conjectured and in part verified between regimes of various superstring theories realized on different backgrounds of the form $M_d\times K$ ( see \cite{duality} and references therein). The existence of such dualities would allow to consider the known superstring theories as local effective formulations on different backgrounds of a unique, yet unknown quantum theory living in a higher dimensional space--time. Among the candidates for this larger quantum theory there is the so called $M$--theory in $(10+1)$--dimensions, whose low energy effective field theory is the well known $11$--dimensional supergravity, and $F$--theory in $(10+2)$--dimensions. Duality, being a mapping between the spectra of two different superstring theories, at the massless level is realized by means of suitable discrete transformations on the background fields, which close a group $G(Z)$. The largest duality which has been conjectured is called $U$--duality and the corresponding action on the massless fields $G(Z)$ is believed to be suitable discrete version of the largest global symmetry group $G$ of the field equations and Bianchi identities of the the underlying low energy supergravity \cite{towhull}. The study of dualities, and in particular of the non--perturbative ones, shed some light on the non--perturbative side of superstring theories. For instance it became clear that the full spectrum of these theories should contain not only $0$ and $1$--dimensional objects (particles and strings), but also $p$--extended objects ($p$--branes) which were not present in the perturbative spectrum, but nonetheless corresponded to solutions of the low energy supergravity. Among them, those coupled with a certain kind of background fields, namely the Ramond--Ramond fields, are of particular interest. The main feature of the Ramond--Ramond (R--R) fields is that they do not couple directly to the fundamental closed string. On the other hand they are mixed with the other background fields which do couple directly to the closed string (Neveu-Schwarz--Neveu-Schwarz fields) through the $U$--duality group $G(Z)$. Therefore, if the $U$--duality is meant to be an exact symmetry of a larger quantum theory, the distinction between R--R and the other fields is to be considered an artifact of perturbative superstring theory and therefore one expects to find in the non--perturbative spectrum of superstring theory objects coupled to R--R fields and thus carrying a R--R charge. Since the R--R fields are in general $(p+1)$ forms, they can minimally couple to $p$--extended objects. A particular kind of these extended objects (namely those preserving a fraction of the original supersymmetry) received a description in the context of perturbative superstring theory on flat space as Dirichelet surfaces \cite{tasi}, that is (roughly speaking) as hypersurfaces on which open strings start and end. The dynamics of these so called $Dp$--branes is determined by the oscillations of the open strings attached to them. In ten dimensions the quantum world may be therefore figuratively represented as containing closed strings free to move in the bulk and to interact among themselves and with dynamical hypersurfaces on which open strings are attached. In eleven dimensions on the other hand, the conjectured $M$--theory is expected to contain $p$--extended objects as well ($Mp$--branes), which are solutions of the low energy supergravity.\par Going back to the suggestive $10$--dimensional picture of the quantum world previously portrayed , one may ask which are the limits of such a representation, i.e. in what physical regime does the framework of superstring theory on Minkowsky background break down? A $D$--brane is a massive object, which means that it interacts with the gravitational field and thus deforms the surrounding space--time. From the superstring point of view, being the quantum fluctuation of the background metric associated with a vibrational $0$--mode of the closed string, this interaction is described through the emission of a closed string from the brane in the surrounding flat space--time and the amplitude of such a process, computed in the framework of perturbative string theory by means, for instance, of boundary state methods, yields the deformation of the Minkowsky metric due to the presence of the extended object. For a single $D$--brane this deformation is localized around the source in a region $R_1$ whose thickness is of the order of the string length $\ell_s$. This is an {\it a posteriori} consistency check that the interaction between the $D$--brane and the closed strings could be correctly computed in the framework of string theory on Minkowsky space, since the closed strings taking part in this interaction ``live'' much longer in the flat region $R_2$, complement of $R_1$ in the whole space, rather than in $R_1$. The situation changes when we consider N coinciding $Dp$--branes. An important feature of these objects is that parallel $D$--branes do not exert any force on each other and therefore may be set to coincide without expense of energy. For this system, the thickness of the curved region $R_1$ around the brane would be proportional to a certain (positive) power of N, and for N large enough, it would contain all the physical information on the interaction between the brane and the closed strings. In the limit $N\rightarrow \infty$ $R_1$ would fill the whole space and the interaction of the strings with the branes could be described only in the framework of a superstring theory realized on a new curved vacuum defined by the near horizon geometry of the branes. Let us look at this scenario from the low energy supergravity point of view.\par The system consisting of a large number $N$ of coinciding $D$ branes, is well described by a solution of the low energy supergravity theory since the curvature in the region $R_1$ is small with respect to $1/{\ell_s}$. In general $p$--brane solutions of supergravity are associated with a $p+1$--form potential $A^{(p+1)}$ to which they couple and whose integral on their world volume gives their {\it electric} charge. (A simple example is that of an elementary particle ($0$--brane) minimally coupled to a vector potential. Its {\it electric} charge is given by the integral of the vector potential along the particle world line.) If $A^{(p+1)}$ is an elementary field of the theory, then the solution is an {\it elementary} $p$--brane, characterized by a singular $p$-hypersurface hidden by an event horizon. To an elementary $p$--brane there corresponds a {\it solitonic} $(D-p-4)$--brane coupled to the $(D-p-3)$--form potential {\it dual} to $A^{(p+1)}$. This {\it bona fide} solution of the theory is non singular. Particular kinds of elementary and solitonic $p$--branes in $D$--dimensions can be found as solutions of a truncated supergravity model consisting just of the metric, the dilaton and the $(p+1)$--form potential \cite{lup,duff,fre} which are described by the following metrics : \begin{eqnarray} \mbox{Elementary solution:}\quad &&\nonumber\\ ds^2\,&=&\,\left(1+\frac{k}{r^{\widetilde{d}}}\right)^{-\frac{4\widetilde{d}}{\Delta(D-2)}}dx^\mu\otimes dx^\nu \eta_{\mu\nu}-\left(1+\frac{k}{r^{\widetilde{d}}}\right)^{\frac{4d}{\Delta(D-2)}}dz^p\otimes dz^q \delta_{p q}\nonumber\\ \mbox{Solitonic solution:}\quad &&\nonumber\\ ds^2\,&=&\,\left(1+\frac{k}{r^d}\right)^{-\frac{4d}{\Delta(D-2)}}dx^\mu\otimes dx^\nu \eta_{\mu\nu}-\left(1+\frac{k}{r^d}\right)^{\frac{4 \widetilde{d}}{\Delta(D-2)}}dz^p\otimes dz^q \delta_{p q} \label{branes} \end{eqnarray} The matrix $\eta_{\mu\nu}$ is the metric on a $(p+1)$--dimensional Minkowsky space--time and has the following signature: $\eta_{\mu\nu}={\rm diag}(+,-,\dots,-)$. These elementary (solitonic) solutions have a flat world volume parametrized by the $d=p+1$ ($\widetilde{d}=D-p-3$) coordinates $x^\mu$, while $z^p$ denote the coordinates along the directions orthogonal to the world volume ($p=1,\dots, d-p-1$). The parameter $\Delta$ is a characteristic quantity which will have the value $4$ in all the cases we shall deal with. Finally $k$ represents the charge of the solution with respect to the corresponding potential. The $r=0$ hypersurface represents the {\it event horizon} of the solution and is a coordinate singularity. An important feature of the above solutions is to have a residual supersymmetry.\par In the $11$--dimensional supergravity (low energy limit of the conjectured $M$--theory) there is an $M2$ (elementary) and an $M5$ (solitonic) brane solution. In type IIB supergravity ($D=10$) of particular interest is the self--dual (since it is coupled to the R--R self dual $4$--form $A^{(4)}$) $3$--brane solution. Their metric is immediately computed from eqs. (\ref{branes}): \begin{eqnarray} \mbox{$D=11$ $M2$--brane: }\quad &&\nonumber\\ ds^2\,&=&\,\left(1+\frac{k_2}{r^6}\right)^{-\frac{2}{3}}dx^\mu\otimes dx^\nu \eta_{\mu\nu}-\left(1+\frac{k_2}{r^6}\right)^{\frac{1}{3}} \left(d^2r +r^2 d^2\Omega_{7}\right) \nonumber\\ \mbox{$D=11$ $M5$--brane: }\quad &&\nonumber\\ ds^2\,&=&\,\left(1+\frac{k_5}{r^3}\right)^{-\frac{1}{3}}dx^\mu\otimes dx^\nu \eta_{\mu\nu}-\left(1+\frac{k_5}{r^3}\right)^{\frac{2}{3}} \left(d^2r +r^2 d^2\Omega_{4}\right) \nonumber\\ \mbox{$D=10$ $3$--brane: }\quad &&\nonumber\\ ds^2\, &=&\, \left(1+\frac{k_3}{r^4}\right)^{-\frac{1}{2}}dx^\mu\otimes dx^\nu \eta_{\mu\nu}-\left(1+\frac{k_3}{r^4}\right)^{\frac{1}{2}} \left(d^2r +r^2 d^2\Omega_{5}\right) \label{M2M5D3} \end{eqnarray} The linear extension of the near horizon region $R_1$ may be characterized by the condition $r\ll R$, $R$ being $k_2^{1/6}$ for the $M2$--brane, $k_5^{1/3}$ for the $M5$--brane and $k_3^{1/4}$ for the $3$--brane. In this region the metrics in eqs. (\ref{M2M5D3}) may be rewritten in the following form: \begin{eqnarray} ds^2\,&=&\,\rho^2 dx^\mu\otimes dx^\nu \eta_{\mu\nu}- (Rw)^2\frac{d\rho^2}{\rho^2}-R^2d\Omega_{(D-p-2)}\nonumber\\ w\,&=&\,\frac{2(D-2)}{\widetilde{d}}\,;\,\rho\,=\,\left(\frac{r}{R}\right)^{\frac{1}{w}} \label{nhorizon} \end{eqnarray} The above metric describes a space--time of the form $AdS_{p+2}\times S^{D-p-2}$ (with an abuse of language we shall denote by Anti de Sitter space a particular compactification of Anti de Sitter space, characterized by the condition $\rho>0$). The parameters $k_2,k_5,k_3$ are related to the charges these solution have with respect to the potential they couple to. Indeed, in suitable units, they represent the flux of the corresponding field strengths $F^{(4)},F^{(7)},F^{(5)}$ through the spheres $S^7, S^4,S^5$ respectively. One may interpret the solutions in eqs. (\ref{M2M5D3}) from a microscopic point of view as bound states describing $N$ coinciding $M2$, $M5$, $D3$ branes respectively, each of them carrying a unit of the charge associated with the potential they couple to. In this picture the parameters $k_i$ ($i=2,5,3$) are proportional to $N$. \par The solutions (\ref{M2M5D3}) describe a space--time continuum which starts from a flat Minkowsky geometry at infinite distance from the extended object, then acquires a non--vanishing curvature at finite $r$ till it forms a ``throat'' of width $R$ near the horizon at $r=0$, with an anti--de Sitter geometry.\par Let us spend few more words about the self--dual $3$--brane solution in type IIB supergravity. If interpreted from the string theory point of view as a bound state of $N$ coinciding $D3$--branes, each carrying unit of R--R charge, its charge with respect to $A^{(4)}$ will be proportional to $k_3=gN\ell_s^4$ ($g$ being the string coupling constant). According to the $D3$--brane interpretation of this solution $R\approx \ell_s (gN)^{1/4}$ and therefore in the limit $N\rightarrow \infty$ the whole space--time has the form $AdS_5\times S^5$. It follows that in this regime a microscopic description of the $D3$--branes would be related to superstring theory realized on this new curved background. Indeed spaces of the form $AdS_{p+2}\times M^{D-p-2}$, where $M^{D-p-2}$ is a suitable Einstein manifold, have been shown to be exact solutions of superstring or M--theory ($D=10,11$ respectively) \cite{kar} and thus possible vacua for these theories (the space $AdS_5\times S^5$ could be interpreted as a non--perturbative vacuum of superstring theory exhibiting a condensation of $D3$--branes). Nevertheless, as previously emphasized, finding the spectrum of physical excitations of a string on a curved background (such as the non--perturbative vacua described above) is a very complicate issue and in general represents an unsolved problem (differently from the case of a conformal $\sigma$--model on a Minkowsky background, in a general curved target space indeed it is no more possible to construct the Hilbert space of quantum states as a Fock space built through the action on a vacuum state of creation operators corresponding to free oscillators). A hint as to the physical content of some of these theories is provided by a powerful duality conjectured by Maldacena almost one year ago. \par In section $2$ I shall give a pedagogical an tentatively self consistent review of this conjecture and its formulation as a duality between a {\it singleton} super--conformal field theory on the boundary of Anti de Sitter space and superstring or M theory in the bulk. In section $3$ I shall focus on the particular case of the $M2$--brane and discuss some recent results in the analysis of the dynamics of branes in Anti de Sitter space. \section{$AdS/CFT$ Duality} Let us focus for the moment on a $10$--dimensional space--time in which the physics is described by type IIB superstring theory in presence of $D3$--branes. The main idea is that, as the number $N$ of coinciding $D3$--branes increases ($N\gg 1$), the physics is described by a new non--perturbative vacuum which is no more the flat Minkowsky one on which the superstring theory is perturbatively defined but has the geometry of $(AdS_5\times K^5)_N$ in which the $N$ $D3$--branes are suitably embedded (the subscript ``$N$'' reminds that the background space--time is the near horizon geometry of the $N$ branes and therefore its ``radius'' is proportional to $N^{1/4}$) . It is reasonable to think that the effective low energy theory around this vacuum has $D3$--branes as fundamental objects instead of strings and describes fluctuations of these extended objects, around their static configuration described by eq. (\ref{M2M5D3}), which have energy much lower than the string scale ($1/\alpha^\prime$). This theory is the effective field theory on the world volume on the $N$ coinciding branes in the limit $\alpha^\prime\rightarrow 0$ (in this limit the interaction of the branes with the bulk through emission of closed strings is suppressed). Intuitively, in much the same way as a field theory of elementary particles and of solitons can be viewed as effective descriptions of a same quantum theory on two different vacua, the theory of superstrings and of branes may be considered as ``dual'' to each other, i.e. the theory defined on the world volume of the $N$ coinciding branes (in which the branes are the fundamental objects) embedded in their near horizon geometry, for small fluctuations, could be viewed as the effective low energy realization of the quantum theory of strings on an $(AdS_5\times K^5)_N$ vacuum. This is the basic idea which has been formalized by Maldacena in his powerful duality conjecture:\par {\it Type IIB superstring theory realized on $(AdS_5\times S^5)_N$ and the low energy effective field theory on the world volume of the $N$ coinciding $D3$--branes are the same quantum theory.} \par The same kind of duality was originally conjectured also between $M$--theory (whose physical content is not known so far) on $(AdS_{p+2}\times S^{9-p})_N$, describing the near horizon geometry of $N$ $Mp$--branes for $p=2,5$, and the low energy effective field theory on the world volume of these branes. In this case the low energy condition on the world volume theory is expressed in terms of the only length scale of the $11$ dimensional supergravity theory, which is the Plank scale $\ell_p$: $\ell_p\rightarrow 0$. This duality is also referred to as {\it holographic} correspondence since it states that the quantum dynamics of fields in a $(p+2)$--dimensional space--time ($AdS_{(p+2)}$) is encoded in a theory defined on a $(p+1)$ dimensional subspace. \par In the original formulation of the conjecture the inner compact space is a $(D-p-2)$--dimensional sphere $S^{(D-p-2)}$. Eventually it was suggested that the duality could be formulated on more general spaces in which the compact Einstein space $K^{(D-p-2)}$ is an homogeneous space of the form $G/H$ \cite{gh}. These spaces are maximally symmetric solutions of the supergravity theory and near horizon geometry of certain $p$--brane solutions of the same theory. Differently form the case in which $K=Sphere$, solutions of the form $AdS_{(p+2)}\times (G/H)$ are in general not maximally supersymmetric, i.e. they preserve less supersymmetry than the original theory. Maldacena's conjecture has been recently further extended to spacetime geometries of the form $AdS\times K$ in which $K$ is an even more general Einstein manifold related to the so called Sasaki monifolds \cite{sasaki}.\par A special role in this duality is played by the Anti de Sitter space--time representing the non--compact factor of this vacuum. Superstring theory (supergravity) realized on $AdS_{(p+2)}\times K^{(D-p-2)}$ is {\it locally} invariant under the general coordinate transformations generated by the space time isometry group: \begin{equation} {\cal G}\,=\,{\cal I}som \left(AdS_{(p+2)}\times K^{(D-p-2)} \right)\,=\, SO(2,p+1)\otimes {\cal I}som \left( K^{(D-p-2)} \right) \label{isom} \end{equation} Since we are dealing with a locally supersymmetric theory, we may include supersymmetry transformations in the previous statement and say that the quantum theory in the bulk is locally invariant with respect to the superextension ${\cal SG}$ of the isometry group in (\ref{isom}), whose structure will be described in more detail in next section. However it is useful to write the explicit form of ${\cal SG}$ in the following way: \begin{eqnarray} {\cal I}som \left( K^{(D-p-2)}\right)&&\rightarrow SO(N)\oplus {\cal K}^\prime \nonumber\\ {\cal SG}\,&=&\, {\cal SC}\oplus {\cal K}^\prime\nonumber\\ {\cal SC}\,&=&\,\cases{Osp(4/{\cal N}) & \mbox{$M2$--brane}\cr Osp(6,2/{\cal N}) & \mbox{$M5$--brane}\cr SU(2,2/{\cal N} ) & \mbox{$D3$--brane}} \label{sg} \end{eqnarray} where ${\cal N}\times$(dimension of the spinors in $(p+2)$ dimensions) gives the number of supercharges preserved by the background solution (${\cal N}$ is the number of {\it Killing spinors} of the supergravity solution and depends on the compact manifold $K$). The group $ {\cal SC}$ is the supersymmetric extension of the group $SO(2,p+1)$, that is it consists of the $SO(2,p+1)$ bosonic generators plus a number ${\cal N}$ of fermionic generators. An important property of the groups $SO(2,p+1)$ and $ {\cal SC}$ is that they act respectively as the {\it conformal} and {\it super conformal} groups in $p+1$ dimensions.\par One of the features of an Anti de Sitter space is to have a boundary. This boundary $\partial AdS_{(p+2)}$ is a $(p+1)$ dimensional locus of points having the property of being stable with respect to the action of the $AdS_{(p+2)}$ isometry group $SO(2,p+1)$ (or the superisometry group ${\cal SC}$ of the Anti de Sitter superspace, in the framework of a supersymmetric theory).\par Parametrizing $AdS_{(p+2)}$ by means of the coordinates $(\rho, x^\mu)$ in terms of which the metric is written in the form (\ref{nhorizon}), the boundary may be characterized as the $(p+1)$--dimensional Minkowsky space $M_{(p+1)}$ spanned by the coordinates $x^\mu$ in the limit $\rho\rightarrow 0$ plus a point at infinity $\rho\rightarrow \infty$. Therefore the Minkowsky part of $\partial AdS_{(p+2)}$ ($M_{(p+1)}$) coincides with the world volume of the $N$ overlapping $p$--branes, which is thus stable under the action of the $SO(2,p+1)$ isometry group.\par The effective low energy theory on the world volume $M_{(p+1)}$ of the $N$ $p$ branes is obtained by expanding the Born--Infeld action $S_{BI}$ on $M_{(p+1)}$ for small oscillations of the branes around the static position at the boundary of $AdS_{(p+2)}$. The action $S_{BI}$ describes a $\sigma$--model on $ M_{(p+1)}$ in which the local fields are the coordinates $X^M(\xi)=(\rho(\xi), x^\mu(\xi);y^m(\xi))$ of the target space--time $AdS_{(p+2)}\times K^{(D-p-2)}$, locally depending on the $p+1$ world volume coordinates $\xi^a$. The main feature of $S_{BI}$ is that it is expressed as the integral over $ M_{(p+1)}$ of the invariant volume element and thus the isometries of the target space are {\it global} symmetries of the theory. In particular, since the Anti de Sitter isometry group $SO(2,p+1)$ acts on $ M_{(p+1)}$ as a conformal group, the theory on the world volume of the $p$--branes on the boundary is a {\it conformal field theory} (CFT). If supersymmetries are taken into account, the theory is expected to be globally invariant under the whole superisometry group ${\cal SG}$, which contains the superconformal group ${\cal SC}$, and thus to be a {\it super conformal} field theory (SCFT).\par The conjectured duality is therefore between superstring ($M$) theory on $AdS_{(p+2)}\times K^{(D-p-2)}$ and a SCFT on the boundary $\partial AdS_{(p+2)}$. \par This conjecture has been subsequently made more precise by Witten \cite{witten} who defined a precise correspondence between quantum states $\Phi (x)$ of the theory in the bulk and superconformal operators ${\cal O}(\xi)$ on the boundary. In particular, according to this picture, the operators ${\cal O}(\xi)$ are interpreted as vertex operators defining the emission of the corresponding state $\Phi (x)$ from the membrane in the bulk. At the heart of this correspondence is the consideration that either the states $\Phi (x)$ or the operators ${\cal O}(\xi)$ must transform in representations of the same group ${\cal G}$ defined in eq. (\ref{isom}) or, more in general, of its superextension ${\cal SG}$. Witten retrieved also a relation between the masses of the states in the bulk and the conformal dimensions of the operators on the boundary. \par As I previously emphasized the whole spectrum of the quantum theory in the bulk is not known, nevertheless the low energy part of it consists of the Kaluza--Klein states of $D$--dimensional supergravity compactified on $AdS_{(p+2)}\times K^{(D-p-2)}$, which have been extensively studied in the eighties \cite{kaluza}. These states are in one to one correspondence with the harmonic functions on the compact space $ K^{(D-p-2)}$ and their mass is of order $m_{KK}\approx 1/R$. They sit in {\it short} supermultiplets of the supergroup ${\cal SG}$ since their masses are related to quantities characteristic of the isometry algebra of $ K^{(D-p-2)}$ and this property protects them, for large enough supersymmetry , from quantum corrections. In the previously outlined correspondence, the Kaluza--Klein states naturally correspond to conformal operators sitting in short representations of the superconformal group ${\cal SC}$ acting on the boundary. They are the {\it chiral} operators. Besides the Kaluza--Klein states there are the string or $M$ theory states which sit on {\it long} supermultiplets of ${\cal SG}$. Their mass is of order $m\approx 1/\ell$ ( $\ell$ being $\ell_s$ or $\ell_p$ in the superstring or $M$--theory case respectively) and are not protected from quantum corrections. These states would naturally correspond to {\it non chiral} operators in the SCFT on the boundary. \par Let us recall that the ``radius'' of the Anti de Sitter space $R$ depends on both $\ell$ and $N$ in the following way: \begin{eqnarray} && R\approx \ell N^\beta \end{eqnarray} where $\beta$ is $1/4$, $1/6$, $1/3$ for the $D3$, $M2$, $M5$ brane respectively. If we send, together with $\ell \rightarrow 0$ (low energy limit on world volume theory) $N$ to infinity in such a way as to keep $R$ fixed, the string or $M$ theory modes will decouple since their mass will grow as $N^\beta$. In this limit the duality can be restated as a mapping between a SCFT on the boundary and tree--level Kaluza--Klein supergravity in the bulk. Since much is known about the latter, Maldacena's duality has so far received several checks from the study of the large $N$ limit of certain SCFT (especially in the case of the low energy SCFT on the world volume of $N$ coinciding $D3$--branes which is known to be a super conformal $SU(N)$ Yang--Mills theory. In this framework stringent check of the duality was carried out for instance in \cite{van}).\par In the correspondence described above between states of the theory in the bulk and conformal operators on the boundary belonging to the same representation of the superconformal group, there is a piece missing. Indeed, among the unitary irreducible representations, besides the long and short supermultiplet, the superconformal group ${\cal SC}$ admits {\it ultrashort} representations ({\it super--singletons}) (in a non--supersymmetric framework, the ultrashort representations of the AdS isometry group $SO(2,p+1)$ are called {\it singletons} and were first found by Dirac \cite{Dirac}). Supersingletons may be characterized as those states which saturate the lowest bound on the energy in order for the representation to be unitary. States in these multiplets don't have a representation in terms of local fields in the bulk since they can be gauged away everywhere, except on the boundary of the Anti de Sitter space, where they admit a field representation . Super--singletons are therefore the natural candidates for describing the elementary fields of the SCFT on the boundary. They don't have a counterpart in the theory defined on the bulk. Superconformal field theories of singletons on $\partial AdS$ have been studied extensively in the eighties \cite{singleton}. An important property of these ultrashort representations is that all the other unitary irreducible representations (UIR) of ${\cal SC}$ (in which for instance all the fields in the bulk transform) may be built from tensor products of two or more singleton states (composite states). This lead, during the early stages of the research on singleton field theories, to the hypothesis that all the excitations of the quantum theory in the bulk could be expressed as composite states of elementary super--singletons living in the boundary (see \cite{duff} and references therein) which was a first kind of {\it holographic } correspondence.\par Already in these early years moreover the super--singleton SCFT was suggested (Nicolai et al. \cite{nicolai}) to describe the dynamics of physical objects embedded on the boundary of $AdS_{(p+2)}$, namely $p$--branes. However this correspondence was never proven rigorously and there are two main differences between the super--singleton SCFT studied in the eighties and those which are expected to describe one side of Maldacena duality. The former were defined on a boundary of $AdS_{(p+2)}$ with topology $S^p\times S^1$, they were (for large enough supersymmetry) free field theories with a mass term and were related to the dynamics of a single {\it spherical } $p$--brane embedded in $\partial AdS_{(p+2)}$. On the other hand SCFTs relevant to Maldacena conjecture (as previously pointed out) are defined on a boundary $\partial AdS_{(p+2)}$ which is given the topology of a $(p+1)$--dimensional Minkowsky space (plus a point at infinity), are interacting (since they have a non--abelian $SU(N)$ gauge invariance) massless field theories and are expected to describe the dynamics of $N$ coinciding $p$--branes set on $\partial AdS_{(p+2)}$. \par Nevertheless, the characterization of the SCFT entering Maldacena duality as a super--singleton theory allowed for a group theoretical analysis of the conjecture \cite{andria}.\par There is however an important feature of the super--singleton theory which is in general not fixed by its invariance with respect to ${\cal SC}$. Indeed we recall that this theory is globally invariant with respect to the whole superisometry group ${\cal SG}$ in eq. (\ref{sg}) which contains, besides ${\cal SC}$, a bosonic subgroup ${\cal K}^\prime$. This group acts as a {\it flavour} group on the singleton theory and the flavour representation of the singletons with respect to ${\cal K}^\prime$ is independent of their superconformal transformation properties and can in general be retrieved only from the $p$--brane dynamics on the boundary. In the maximally supersymmetric case $K^{(D-p-2)}=S^{(D-p-2)}$ the whole ${\cal I}som \left(S^{(D-p-2)}\right)=SO(D-p-1)$ enters the superconformal group ${\cal SC}$ in eqs. (\ref{sg}) and the group ${\cal K}^\prime$ is trivial, therefore all the transformation properties of the singletons with respect to ${\cal SG}$ are characterized by the ${\cal SC}$ quantum numbers. Verifying Maldacena conjecture in the most general case requires the knowledge of the flavour representation of the singletons with respect to ${\cal K}^\prime$. Indeed since all the conformal operators of the theory are constructed as composite structures of the singleton fields, their ${\cal K}^\prime$ quantum numbers will depend on the singleton flavour representation. Verifying the matching between the ${\cal K}^\prime$ quantum numbers of the boundary operators with those of the bulk states would be a new kind of check of Maldacena conjecture. To this end it is crucial to develop a procedure for retrieving the supersingleton action from the quantization of a $N$ coinciding $p$--branes set on the boundary of $ AdS_{(p+2)}$. For $N>1$ a classical action on the world volume of the overlapping branes would be a non--abelian Born--Infeld action, which is not known at present. However, if our aim is just to determine the ${\cal K}^\prime$ flavour representation of the singletons, this information could be inferred from the study of a single $p$--brane on the boundary. Indeed one may consider a system of $N+1$ parallel $p$--branes, of which the first $N\gg 1$ are coinciding and set on the boundary of $AdS_{(p+2)}\times K^{(D-p-2)}$ while the remaining one ({\it probe} brane) is set at a distance $r$ from them. In principle then one could quantize (for small fluctuations) the world volume Born--Infeld action on the probe brane in the limit in which the latter is sent to coincide with the other $N$ ($r\rightarrow 0$) (superconformal limit) and retrieve an the superconformal action of free (abelian) massless singletons, together with their ${\cal K}^\prime$ representation.\par Quantizing for small oscillations the B--I action on the world volume of a brane embedded in a curved space is a difficult problem because of technical difficulties related to the gauge fixing of the local invariances of the theory (in particular the fermionic $\kappa$ symmetry needed in order for the world volume theory to be supersymmetric). A general method for retrieving the supersingleton action from the world volume theory of a probe brane on the boundary of an Anti de Sitter space was defined in \cite{osp}, and applied to the case of an $M2$--brane on the boundary of $AdS_4\times S^7$. This method relied on a particular parametrization of the Anti de Sitter superspace defined by using a {\it supersolvable} subalgebra of the superisometry group. The supersolvable parametrization is the super extension of the {\it solvable} parametrization of the Anti de Sitter space defined in \cite{gh}, which is related to the coordinates $(\rho, x^\mu)$ in terms of which the metric has the expression (\ref{nhorizon}). The main goal of the supersolvable parametrization is to allow to write the B--I action which is $\kappa$--symmetry fixed from the very beginning, that is which depends only on the physical fermionic fields. This $\kappa$--gauge fixing turns out to be equivalent to fixing the {\it Killing spinor} gauge defined in \cite{kall}.\par Solvable and Supersolvable parametrizations defined in \cite{gh} and \cite{osp} revealed to be a powerful algebraic tool for describing brane dynamics in Anti de Sitter spaces. They have been applied eventually to the $D3$ brane in $AdS_5\times S^5$ \cite{pesando}.\par The final goal however is to apply this method to a background of the form $AdS_{(p+2)}\times K^{(D-p-2)}$ where $K$ is an homogeneous manifold of the form $G/H$ or related to a more general Sasaki manifold (${\cal N}=2$ Killing spinors). In these cases indeed, as previously emphasized, the $AdS/CFT$ correspondence from a group theoretical point of view becomes non--trivial because the ${\cal K}^\prime$ flavour representation of the singletons enters the game on the SCFT side, as an input independent of the superconformal invariance of the theory. This analysis is still work in progress. \section{Supermembrane in Anti de Sitter space.} In the present section I shall recall the main facts about Anti de Sitter spaces and review some recent results in the study of the $M2$--brane dynamics in $AdS_4\times S^7$.\par An $n$--dimensional Anti de Sitter space--time may be described as an hypersurface embedded in $\relax{\rm I\kern-.18em R}^{2,n-1}$. Let $X^\Lambda=(X^{0},X^{\overline{0}}, X^1,\dots, X^{n-1})$ denote a coordinate basis for $\relax{\rm I\kern-.18em R}^{2,n-1}$ with respect to which the flat metric has the form: $\eta_{ \Lambda\Sigma}={\rm diag}(+,+,-,\dots,-)$. $AdS_n$ is defined as the following locus of points: \begin{eqnarray} \eta_{ \Lambda\Sigma}X^\Lambda X^\Sigma\,&=&\, 1 \label{hyper} \end{eqnarray} (for simplicity we have set the ``radius'' of $AdS_n$ equal to $1$). This condition is invariant with respect to the group $SO(2,n-1)$ acting on $X^\Lambda$, which is the isometry group of $AdS_n$. Defining the light cone coordinates $X^{\pm}=X^{\overline{0}}\pm X^{n-1}$ and the coordinates $x^\mu$ by setting $X^\mu=X^+x^\mu $ for $\mu=0,1,\dots,n-2$, the condition (\ref{hyper}) may be rewritten in the following way: \begin{eqnarray} X^+X^-+(X^{+})^2\eta_{ \mu\nu}x^\mu x^\nu \,&=&\, 1 \label{hyper2} \end{eqnarray} Let us restrict ourselves to the branch $X^+\ge 0$. In the limit $X^+\gg 1$ the above condition reduces to: \begin{eqnarray} X^{+}\eta_{ \mu\nu}x^\mu x^\nu \,&=&\, -X^- \end{eqnarray} which defines an $(n-1)$--dimensional Minkowsky space $M_{(n-1)}$ spanned by the coordinates $x^\mu$, which is stable with respect to a scaling $X^\Lambda\rightarrow \lambda X^\Lambda$. Its isometry group is $ISO(1,n-2)\subset SO(2,n-1)$. Moreover if we add to it the point P at $X^+=0$ defining $\widetilde{M}_{(n-1)}=M_{(n-1)} \cup P(X^+=0)$, this compact locus of points defines the boundary of $AdS_n$ which is stable with respect to the whole group $SO(2,n-1)$. Setting $\rho=X^+$ and computing the induced metric on the hypersurface (\ref{hyper2}) in terms of the coordinates $(\rho,x^\mu)$ one obtains a generalization of the Bertotti--Robinson metric: \begin{eqnarray} ds^2=\rho^2 dx^\mu \otimes dx^\nu \eta_{\mu\nu} -\frac{d\rho^2}{\rho^2} \label{BR} \end{eqnarray} The Anti de Sitter space may be alternatively written as an homogeneous manifold in the following way: \begin{equation} AdS_n=\frac{SO(2,n-1)}{SO(1,n-1)} \label{coset} \end{equation} In order to compute the non--linear action of the isometry group $SO(2,n-1)$ on the coordinates $(\rho, x^\mu)$, we shall characterize them as coordinates in the {\it solvable} representation of $AdS_n$ \cite{gh}. Using the formalism of coset manifolds, one can associate with each point Q in $AdS_n$ a {\it coset representative} $L(Q)$ so that the non linear action of the isometry group $SO(2,n-1)$ on the point Q may be represented in the following way: \begin{eqnarray} \forall g\in SO(2,n-1)&&\,\,\, g\cdot L(Q)\,=\, L(Q^\prime(g,Q))\cdot h(g,Q)\nonumber\\ h(g,Q)\in SO(1,n-1) \label{g} \end{eqnarray} In order to define a parametrization of the manifold, or equivalently $L(Q)$, one may use the following algebra decomposition: \begin{eqnarray} SO(2,n-1)\,&=&\, SO(1,n-1)\oplus Solv \label{iwa} \end{eqnarray} where $Solv$ is an $n$--dimensional {\it solvable}\footnote{A solvable Lie algebra $Solv$ is a Lie algebra whose $k^{th}$ Lie derivative ${\cal D}^k(Solv)$ vanishes for some finite $k$. The $k^{th}$ Lie derivative may be defined by induction in $k$: ${\cal D}^{k+1}(Solv)=\left[{\cal D}^{k}(Solv),{\cal D}^{k}(Solv)\right]$; ${\cal D}^{1}(Solv)=\left[Solv,Solv\right]$ } subalgebra of $SO(2,n-1)$. As a consequence of eq. (\ref{iwa}) the Anti de Sitter space is isomorphic to a group manifold generated by $Solv$ and may be globally identified with it. In this description the local coordinates on $AdS_n$ are therefore parameters of $Solv$. The $Solv$ algebra is constructed by considering the only non--compact Cartan generator $D$ in the coset (\ref{coset}) and diagonalizing its adjoint action on the algebra $SO(2,n-1)$. This defines a {\it grading} on $SO(2,n-1)$ which is decomposed into the direct sum of eigenspaces of $Adj_D$, namely $g_{(0)},g_{(\pm1)}$ corresponding to the eigenvalues $0$, $\pm1$ respectively. $Solv$ can be constructed as $\{D\}\oplus g_{(-1)}$, where $g_{(0)}=\{D\}\oplus SO(1,n-2)$ while $g_{(-1)}=\{T_\mu\}$ consists of the $n-1$ {\it shift operators} corresponding to all the roots having negative value on $D$. The subalgebra $g_{(-1)}$ is an abelian subalgebra of $SO(2,n-1)$ since there is no generator with grading $-2$ with respect to $Adj_D$. Associating with the generator $D$ the parameter ${\rm log}(\rho)$ and with the generators $T_\mu$ the parameters $x^\mu$, the coset representative can be defined as an element of the solvable group ${\rm Exp}(Solv)$ as follows: \begin{equation} L(\rho,x^\mu)\,=\,{\rm Exp}\left({\rm log}(\rho) D+ \sum_{\mu=0}^{n-1}x^\mu T_\mu\right) \label{cosetrep} \end{equation} Constructing the vielbeins in this parametrization and then computing the metric on the manifold, one obtains the expression in eq. (\ref{BR}). From eqs. (\ref{cosetrep}) and (\ref{g}) one may therefore deduce the action of $SO(2,n-1)$ on the coordinates $(\rho, x^\mu)$. \par In this parametrization Anti de Sitter space is represented, roughly speaking, by means of Minkowsky ``slices'' $M_{(n-1)}$ spanned by the coordinate $x^\mu$ and fibrated along the transverse coordinate $\rho \in \relax{\rm I\kern-.18em R}^+$.One may define an alternative topology on $\partial AdS_{n}$, by constructing it as the limit of $M_{(n-1)}$ as $\rho\rightarrow 0$ plus a point at $\rho\rightarrow \infty$ \cite{osp}. Thinking of $AdS_n$ as the near horizon geometry on $N$ coinciding $(n-2)$ branes set on the $\rho=0$ chart of the boundary, the {\it probe} brane introduced in last section, can be defined as a brane whose wold volume coincides with the $M_{(n-1)}$ Minkowsky ``slice'' at a distance $\rho$ from the remaining $N$ branes.\par It is possible to check that the action of the group $SO(2,n-1)$ on a generic hypersurface $M_{(n-1)}$ at $\rho\neq 0$ closes a ``soft'' conformal algebra, that is a conformal algebra whose structure constants depend on the coordinate $\rho$ ( the {\it broken conformal transformations} of \cite{malda}). In the limit $\rho\rightarrow 0$ (on the boundary) the dependence on $\rho$ of the conformal algebra drops out and the group $SO(2,n-1)$ acts on $M_{(n-1)}$ as the usual conformal group. Now we can interpret the generators of $SO(2,n-1)$ described above, from the world volume theory point of view, in the following way: \begin{itemize} \item $D$ the generator of the {\it dilations} \item $T_\mu$ the $n-1$ {\it translations} on $M_{(n-1)}$ \item $K_\mu$ in $g_{(+1)}$ as the $n-1$ {\it special conformal transformations} \item $SO(1,n-2)=g_{(0)}/\{D\}$ the {\it Lorentz transformations} on $M_{(n-1)}$ \end{itemize} In what follows I shall outline, without entering the details of the calculations, the main conceptual steps in the procedure introduced in \cite{osp} for computing the supersingleton action from the quantum fluctuations of a probe $p$--brane around the boundary of an Anti de Sitter space. Let us consider for simplicity the special case of an $M2$ probe brane in $AdS_4\times S^7$. As previously pointed out this space--time is a maximally supersymmetric solution of the $11$--dimensional supergravity. With it we may associate a superspace $AdS^{(4/8)}\otimes S^7$ spanned by $10$ bosonic coorsinates $X^M$ ($M=0,\dots, 9$) and $32$ fermionic coordinates $\Theta^A_\alpha$ where $A=1,\dots, {\cal N}=8$ and $\alpha=1,\dots,4$ indicizes a $4$--dimensional spinor. $X^M$ consist in the four coordinates $(\rho, x^\mu)$ of $AdS_4$ and the seven coordinates $y^m$ of $S^7$. The superspace has the form: \begin{eqnarray} AdS^{(4/8)}\otimes S^7\,&=&\, \frac{Osp(4/8)}{SO(1,3)\times SO(8)}\otimes S^7 \label{sspace} \end{eqnarray} The theory on the world volume $M_{(3)}$ of the $M2$ probe brane is described by a Born--Infeld action $S_{BI}$ which, as explained in the previous section, is a $\sigma$--model defined on $M_{(3)}$ having the background superspace $AdS^{(4/8)}\otimes S^7$ as target space. The local fields in this action are therefore $X^M(\xi)$ (bosons) and $\Theta^A_\alpha(\xi)$ (fermions), $\xi^a$ being the $3$ world volume coordinates. Using the conventions of \cite{osp} $S_{BI}$ may be written (in suitable units) in the following form: \begin{eqnarray} S_{BI}\,&=&\, 2\int_{M_{(3)}}\sqrt{-{\rm det}(h_{ab})}+(4!)^2\int_{M_{(3)}} A^{(3)} \label{biaction} \end{eqnarray} where $h_{ab}(X^M(\xi),\Theta^A_\alpha(\xi))$ is the induced metric on the world volume and the second term is the Wess--Zumino term describing the minimal coupling of the brane with the $3$--form of the $11$--dimensional supergravity.\par Let us analyze the symmetries of this action. As previously stated, its {\it global} symmetries are the target space superisometries which close the group ${\cal SG}=Osp(4/8)$. Its {\it local} symmetries are, on the other hand, the {\it diffeomorphisms} on the world volume ($\xi\rightarrow \xi^\prime(\xi)$) and the fermionic $\kappa$--symmetry. The former implies that the only physical degrees of freedom are the oscillations of the membrane in the directions perpendicular to its world volume ($8$ bosonic d.o.f.). The latter is a symmetry of the action with respect to local transformations acting only on half the components of the fermion fields. These components are defined by a projector of the form ${\cal P}^+=(1+\Gamma)/2$, where $\Gamma$ is a $4\times 4$ matrix depending on the background fields. The invariance of $S_{BI}$ with respect to this local transformation requires the background fields to be a solution of the $11$ dimensional supergravity, which is indeed the case.The consequence of $\kappa$ symmetry is that the $\Theta^+={\cal P}^+\Theta$ components of the fermionic fields can be gauged away, leaving $16$ physical fermionic degrees of freedom $\Theta^-$ which match {\it on--shell} the $8$ bosonic ones. The super--singletons indeed belong to $8$ supermultiplets consisting of a boson and a Majorana fermion each. These fields will be retrieved as quantum fluctuations around a suitable solution of the $8$ fermionic and $8$ bosonic physical degrees of freedom of the brane. The existence of $\kappa$ symmetry is therefore required in order for the theory on the world volume to be on--shell supersymmetric.\par A particular solution of the theory is the static one: \begin{eqnarray} x^\mu(\xi)\, &=&\, \xi^a \,\, \mu,\, a\,=\, 0,1,2\nonumber\\ \rho(\xi)\,&\equiv&\, \overline{\rho}\,=\, const\nonumber\\ \partial_a y^m\, &=&\,0\nonumber\\ \Theta^A_\alpha\, &=&\,0 \label{static} \end{eqnarray} In order to retrieve the supersingleton action the strategy is to fix the local symmetries of the theory and expand the action (\ref{biaction}) for small fluctuations of the physical d.o.f. around the solution (\ref{static}) and then taking the boundary limit $\overline{\rho}\rightarrow 0$ where the full superconformal symmetry is restored. The local diffeomorphisms are fixed by fixing the coordinates $x^\mu$ parallel to the brane to the solution (\ref{static}). Fixing the $\kappa$ symmetry is a more complicate issue. In principle one should compute the components $\Theta^\pm$ on the static solution, set the unphysical components $\Theta^+$ to zero in the action and allow the physical ones to fluctuate. This is difficult to implement on the action (\ref{biaction}) initially expressed in terms of the whole spinors $\Theta_\alpha^A$, since on the Anti de Sitter background, this dependence is rather complicate and involves higher powers of the $\Theta$ fields.\par This problem has been circumvented in \cite{osp} by using the supersolvable parametrization of the superspace (\ref{sspace}) which consists in redefining the target space of our $\sigma$ model to be a subspace of the whole superspace which is a supermanifold generated by a {\it supersolvable} algebra ($SSolv$) (times of course the sphere $S^7$spanned by $y^m$).\footnote{The definition of a supersolvable Lie algebra is the same as that of a solvable algebra but with the supercommutator substituted to the commutator} We start from a decomposition analogous to that in (\ref{iwa}): \begin{eqnarray} Osp(4/8)\,&=&\, \left[SO(1,3)\oplus SO(8)\oplus {\cal Q}\right]\oplus Ssolv \end{eqnarray} The above decomposition is obtained by performing a grading of the superalgebra with respect to $Adj_D$: \begin{eqnarray} Osp(4/8)\,&&\rightarrow g_{(-1)}\oplus sg_{(-1/2)}\oplus g_{(0)}\oplus sg_{(1/2)}\oplus g_{(1)}\nonumber\\ {\cal Q}\,&=&\,sg_{(1/2)}\nonumber\\ {\cal S}\,&=&\,sg_{(-1/2)}\nonumber\\ SO(1,3)\oplus SO(8) \oplus \{D\}\,&=&\,g_{(0)}\nonumber\\ SSolv\,&=&\,\{D\}\oplus g_{(-1)}\oplus sg_{(-1/2)}\,=\, Solv\oplus sg_{(-1/2)} \end{eqnarray} the spaces $g_{(\pm 1)}$ are the same as those defined previously for the Anti de Sitter space. The subspaces $sg_{(\pm 1/2)}$ are the eigenspaces with of the fermionic generators corresponding to the eigenvalues $\pm 1/2$ of $Adj_D$: on the world volume theory the ${\cal S}$ operators ($16$ components) generate the {\it special superconformal transformations}, while the ${\cal Q}$ operators ($16$ components) generate the {\it supersymmetry transformations}. Differently from the solvable representation of Anti de Sitter space, the supergroup ${\rm Exp} (SSolv)$ does not coincide with the original superspace since the generators ${\cal Q}$ are modded out. Nevertheless it can be shown that $Adj_D$ on the fermionic generators is represented by an operator $\widetilde{\Gamma}$ which coincides with the $\kappa$ symmetry operator $\Gamma$ on the static solution (\ref{static}) (it is expressed as the product of the gamma matrices along the directions of the world volume). In other words, the $\kappa$--symmetry projectors ${\cal P}^\pm$ computed on the solution (\ref{static}) coincide with the projectors of the supersymmetry generators into the eigenspaces $sg_{(\pm 1/2)}$ respectively. Therefore the fermionic coordinates parametrizing $SSolv$ are already the $\kappa$ gauge fixed ones. In this parametrization indeed the coordinates are $(\rho, x^\mu, y^m ; \Theta^-)$. The B--I action defined on this supersolvable target space therefore is $\kappa$ gauge fixed from the very beginning and much simpler to compute. \par Expanding this action around the static solution for small fluctuations of the physical fields, rescaling the latter by suitable powers of $\overline{\rho}$, taking the order $\alpha^{\prime 0}$ of the action and sending $\overline{\rho}\rightarrow 0$ it was possible to retrieve the supersingleton action on the boundary as a free superconformal field theory describing $8$ massless bosons and $8$ massless Majorana fermions. \section{Conclusions} The aim of the present talk is on one hand to give a tentatively self--contained and hopefully elementary introduction to Maldacena's conjecture and on the other hand to frame within a discussion on the intense research devoted to verify this conjecture and to possibly extend it, some recent results in the study of brane dynamics in certain Anti de Sitter spaces. I emphasized how some aspects of the super--singleton CFT on one side of this duality could be inferred only from the dynamics of $p$--branes on the boundary of a $AdS_{(p+2)}\times K^{(D-p-2)}$ space--time for a general compact Einstein manifold $K$. The knowledge of these aspects would provide the basis for a new stringent check of the conjecture. So far a method for constructing the super--singleton action on the world volume of a probe brane at the boundary of $AdS_{(p+2)}\times S^{(D-p-2)}$ has been defined by means of a {\it normal coordinate expansion} (small quantum fluctuations) of the B--I action on the probe brane around the static configuration in which the latter is set to lie on the boundary of the Anti de Sitter space. An extension of this method to more general internal spaces $K$ is still work in progress.
\section{Introduction} \label{sec:intro} In his recent book {\it Consilience: The Unity of Knowledge}, biologist E. O. Wilson reintroduced the word ``consilience'' into the English language. William Whewell wrote in 1840 that ``The Consilience of Induction takes place when an Induction, obtained from one class of facts, coincides with an Induction, obtained from another different class. This Consilience is a test of the truth of the Theory in which it occurs~\cite{EOW}.'' The cuprate high temperature superconductors are complex enough that it is appropriate to apply this truth principle to sort out various models. In this brief overview, I focus on {\it some} of the theories which have at least a chance of surviving. I will argue that recent years have seen a consilience, a coming-together, of theories of high temperature superconductivity. This lecture was presented in Hanoi at the {\it International Workshop on Superconductivity, Magneto-Resistive Materials, and Strongly Correlated Quantum Systems}. Vietnam is a country with a long tradition of Buddhism. It is therefore fitting to recall the second of Buddhism's Four Noble Truths: Attachment leads to suffering. It is important to not become too attached to any one theory of high-temperature superconductivity! Some are completely wrong; others mix strong and weak features, and none are as yet definitive. \section{Key Experiments} \label{sec:key} Recent experimental advances show that the various high temperature superconductors share much in common. A good overview of the cuprate materials and their key experimental properties can be found in a recent review by Maple~\cite{Maple}. Experiments now agree that there is d$_{x^2-y^2}$ superconducting order~\cite{d-wave}, at least in the hole-doped compounds. One of the most intriguing phenomena is the existence of a pseudogap in the density of states at temperatures {\it above} the superconducting transition temperature $T_c$. Many experiments show, or are at least consistent with, the existence of the pseudogap~\cite{Service,Deutscher,Corson,Markiewicz,Kusko}. Four other key experimental results are: 1. Rough Electron-Hole Symmetry: As emphasized by Maple~\cite{Maple}, the behavior of the electron doped materials, such as Nd$_{2-x}$Ce$_x$CuO$_{4-y}$, is qualitatively similar to that of the hole-doped compounds such as La$_{2-x}$Sr$_x$CuO$_{4-y}$. In particular, both are antiferromagnetic insulators at small $x$ which become superconducting at larger values of $x$. Whether or not the electron doped materials are d-wave superconductors is a crucial question. If they are s-wave, much of the subsequent discussion in this article is falsified, or at the very least must be reworked. 2. Angle Resolved Photoemission (ARPES): Although ARPES experiments continue to contradict each other~\cite{Norman,Ino,Chuang}, some features stand out. Nesting of the Fermi surface seems to have been observed~\cite{Dessau}. And one insulating antiferromagnetic parent compound shows a very interesting electron dispersion~\cite{Ronning}, similar to that of a d-wave superconductor, despite the absence of superconductivity, see Sec.~ \ref{sec:mean-field} below. 3. Neutron Scattering: Neutrons are sensitive to both static and fluctuating magnetic moments. A useful summary is presented by Mason~\cite{Mason}. Incommensurate magnetic fluctuations have been observed for some time in doped 214 compounds~\cite{Aeppli,Wells,Lee,Wakimoto}, and the deviation of the peak wavevectors from the N\'eel ordering wavevector $\vec{Q} = (\pi, \pi)$ is proportional to the doping. The recent observation of similar incommensurate peaks in the 123 material~\cite{Mook} suggests that this behavior is common, and perhaps universal, to the cuprates. 4. Phase Separation and Stripes: Although phase separation into {\it static} stripes may occur only in some of the cuprates~\cite{Tranquada,Julien}, and at special dopings such as $x = 1/8$, there is mounting evidence for ubiquitous, but slowly shifting stripes, in particular from $^{63}$Cu NQR measurements~\cite{Hunt}. Even when there is no true long-range stripe order, the stripe dynamics may be slow enough that the stripes can be treated for most purposes as static. \section{Hubbard and Related Models} \label{sec:models} Anderson's early insight was that the Hubbard model is a good starting point for theories of cuprate superconductivity~\cite{PWAScience}. The one-band model, \begin{equation} H = -t~ \sum_{<{\bf x},{\bf y}>}~ (c^{\dagger \alpha}_{\bf x} c_{{\bf y} \alpha} + H.c.) + U~ \sum_{\bf x}~ (c^{\dagger \alpha}_{\bf x} c_{{\bf x} \alpha} - 1)^2\ , \end{equation} can be justified starting from the more complete three-band model which describes the p$_{x,y}$ oxygen orbitals as well as the d$_{x^2 - y^2}$ orbitals of the copper atoms~\cite{Auerbach}. For simplicity only nearest-neighbor hopping, and on-site Coulomb repulsion, appear in the above Hamiltonian; however, next-nearest neighbor hopping can be important. Furthermore, the assumption that the inverse-square Coulomb interaction is screened down to a pure on-site repulsion certainly breaks down if there is phase separation (see Sec. ~\ref{sec:phase-sep}). At half-filling, and on a square lattice, nesting instabilities open up a charge gap and the system becomes an antiferromagnetic insulator. The low energy spin degrees of freedom can be described in terms of the electron operators as: \begin{equation} \vec{S}_{\bf x} = {\frac{1}{2}} c_{\bf x}^{\dag \alpha}~ \vec{\sigma}_\alpha^\beta~ c_{{\bf x} \beta} \label{fermi-spin} \end{equation} subject to the constraint \begin{equation} n_{\bf x} \equiv c_{\bf x}^{\dag \alpha} c_{{\bf x} \alpha} = 1 \ . \end{equation} The representation of the spins in terms of the underlying fermions is not as economical as the representation of the spins themselves, since the spin operators of Eq. ~\ref{fermi-spin} are invariant under $U(1)$ gauge transformations at each lattice point~\cite{Baskaran,AM} \begin{eqnarray} c_{{\bf x} \alpha} &\rightarrow& e^{i \theta_{\bf x}}~ c_{{\bf x} \alpha} \nonumber \\ c^{\dag \alpha}_{\bf x} &\rightarrow& e^{-i \theta_{\bf x}}~ c^{\dag \alpha}_{\bf x} \label{xU(1)} \end{eqnarray} because the local phase rotation $\theta_{\bf x}$ cancels out. We denote this infinite gauge symmetry $U(1)^\infty_x$. The physical meaning of this local symmetry is simply that the charge degrees of freedom are frozen out and play no role in the insulating magnet. The underlying fermions have both charge and spin degrees of freedom, but as the number of fermions on each site is fixed to be one, only the spin degree of freedom is active. Conversely, in the $U \rightarrow 0$ limit, the Hubbard model reduces to a non-interacting tight-binding model \begin{equation} H = -t~ \sum_{<{\bf x},{\bf y}>} c_{\bf x}^{\dag \alpha} c_{{\bf y} \alpha} + H.c. \end{equation} which is clearly not invariant under the $U(1)^\infty_x$ transformations, Eq. ~\ref{xU(1)}. Distinct $U(1)$ rotation angles at neighboring sites, $\theta_x$ and $\theta_y$ do not cancel out. Physically this just means that the electron number is not conserved as the electrons hop from site to site. However, in momentum space, \begin{equation} H = \sum_{{\bf k}} \epsilon_{{\bf k}}~ c^{\dag \alpha}_{{\bf k}} c_{{\bf k} \alpha}; \ \ \ \ \epsilon_{{\bf k}} = -2 t [\cos(k_x) + \cos(k_y)]\ . \end{equation} The Hamiltonian is instead invariant under a different infinity of $U(1)$ rotations~\cite{Haldane}, now at each point in momentum space, rather than position space: \begin{eqnarray} c_{{\bf k} \alpha} &\rightarrow& e^{i \theta_{\bf k}}~ c_{{\bf k} \alpha}\ . \nonumber \\ c^{\dag \alpha}_{{\bf k}} &\rightarrow& e^{-i \theta_{\bf k}}~ c^{\dag \alpha}_{{\bf k}}\ . \label{kU(1)} \end{eqnarray} This $U(1)^\infty_k$ symmetry, like the $U(1)^\infty_x$ symmetry explored above, has a clear physical interpretation: in the absence of interactions and disorder, momentum is a good quantum number for the single-particle states, as these are infinitely long-lived. In summary, \begin{eqnarray} U(1)^\infty_x &\leftrightarrow& {\rm insulating~ solid} \nonumber \\ U(1)^\infty_k &\leftrightarrow& {\rm conducting~ liquid.} \end{eqnarray} The real cuprates lie somewhere between these two extreme limits, and it is the tension between these two limits that leads to much interesting physics. Further progress can be made by noting that $U \gg t$; hence, double-occupied sites are energetically disfavored in hole-doped compounds. We can enforce this reduction of the on-site Hilbert space from 4 states to 3 by employing slave bosons to represent the holes and write the t-J model~\cite{Auerbach} as follows: \begin{equation} H = -t~ \sum_{<{\bf x},{\bf y}>}~ (c^{\dagger \alpha}_{\bf x} b_{\bf x} c_{{\bf y} \alpha} b^\dagger_{\bf y} + H.c.) + J~ \sum_{<{\bf x},{\bf y}>}~ (\vec{S}_{\bf x} \cdot \vec{S}_{\bf y} - \frac{1}{4} n_{\bf x} n_{\bf y}) \ . \end{equation} Every time an electron $c$ hops from site ${\bf x}$ to ${\bf y}$, a hole $b$ hops in the reverse direction. The holonomic constraint $c^{\dagger \alpha}_{\bf x} c_{{\bf x} \alpha} + b^\dagger_{\bf x} b_{\bf x} = 1$ then forces the total electron number on each site to be equal to zero or one, but not two. There are many approximate solutions of the t-J model. In the following sections I describe several different starting points which, surprisingly, lead to rather similar conclusions. \section{Mean-Field and Gauge Theories of the Undoped Antiferromagnet} \label{sec:mean-field} The first starting point I consider is a systematic solution of the $t-J$ model by means of a $1/N$ expansion~\cite{AM,MA}. The basic idea is to generalize the usual two types of spin, up and down, to $N$ types, and solve the resulting $SU(N)$ t-J model in the $N \rightarrow \infty$ limit. Fluctuations disappear in the $N \rightarrow \infty$ limit and complex-valued mean-fields along the bonds acquire expectation values: \begin{equation} \chi_{{\bf x} {\bf y}} = \frac{J}{N} c^{\dagger \alpha}_{\bf x} c_{{\bf y} \alpha} \ . \end{equation} An additional set of scalar fields, $\phi_{\bf x}(t)$, are introduced as Lagrange multipliers to enforce the holonomic constraint on the occupancy. Under the local $U(1)$ gauge transformations, Eq. ~\ref{xU(1)}, the $\chi$ fields transform like the spatial link variables of a compact lattice gauge theory, while the $\phi$ fields transform as the time-component of the gauge fields: \begin{eqnarray} \chi_{{\bf x} {\bf y}}(t) &\rightarrow& e^{i[\theta_{\bf y}(t) - \theta_{\bf x}(t)]}~ \chi_{{\bf x} {\bf y}}(t) \nonumber \\ \phi_{\bf x}(t) &\rightarrow& \phi_{\bf x}(t) + \frac{d \theta_{\bf x}(t)}{dt}\ . \end{eqnarray} One annoying problem with the large-N limit is that, close to half-filling, the lowest-energy solution is dimerized: $\chi_{{\bf x} {\bf y}}$ is non-zero only on disconnected bonds. This solution, which is unphysical, can be eliminated by the introduction of biquadratic spin-spin interactions~\cite{MA} which do nothing to the original $SU(2)$ model (as they reduce to the usual bilinear exchange at $N=2$) but which suppress dimerization for $N > 2$. Furthermore, instantons do not induce dimerization as they do in the bosonic $SU(N)$ and $Sp(N)$ formulations~\cite{absence}. This is clearly demonstrated by an exact solution which shows no dimerization~\cite{C-break}. At half-filling the ground state is the $\pi$-flux phase~\cite{AM,Kotliar}. The flux is defined by the gauge invariant plaquette operator $\langle \chi_{12} \chi_{23} \chi_{34} \chi_{41} \rangle < 0$ with the $\chi$ fields oriented as shown in Fig. ~\ref{fig:flux}. Rokhsar pointed out~\cite{Rokhsar} a simple way to see that non-zero flux is energetically favored: it lowers the kinetic energy of the fermions on the lattice, an effect which is particularly easy to see for the four-site problem, see Fig. ~\ref{fig:flux}. \begin{figure} \epsfig{figure=flux.eps,height=2.5in} \caption{Orientation of the $\chi_{{\bf x} {\bf y}}$ fields in the $\pi$-flux phase. Since $-\pi$ flux is equivalent to $\pi$ flux, this phase has the full translational symmetry of the underlying lattice and respects time-reversal symmetry. By considering a single plaquette of just four sites, it is clear that $\pm \pi$ flux is energetically favored as the filled electronic states have lower net energy (solid dots) compared to zero flux (open dots). \label{fig:flux}} \end{figure} The spectrum is markedly different from that of a tight-binding model as now the fermion dispersion has the form: \begin{equation} \epsilon(\vec{k}) \propto \pm \sqrt{\cos^2(k_x) + \cos^2(k_y)}\ . \end{equation} The negative energy states are filled, and the nesting instability has been removed by the appearance of a pseudogap everywhere except at the discrete points $\vec{k} = (\pm \pi/2, \pm \pi/2)$ where the density of states vanishes linearly as shown in Fig. ~\ref{fig:dispersion}. \begin{figure} \epsfig{figure=disp.eps,height=2.5in} \caption{Dispersion of the flux phase. There is a pseudo-gap everywhere except at the nodal points $\vec{k} = (\pm \pi/2, \pm \pi/2)$. These points are paired and the dispersion is linear (cone) like that of massless Dirac fermions. \label{fig:dispersion}} \end{figure} It is remarkable that precisely this dispersion is seen in ARPES experiments~\cite{Ronning} on the insulating parent compound Ca$_2$CuO$_2$Cl$_2$, an observation emphasized by Laughlin~\cite{Laughlin}. Because the mean-fields $\chi$ and $\phi$ are singlets under $SU(N)$ spin rotations, long-range spin-order is impossible in the $N \rightarrow \infty$ limit. To recover N\'eel order requires the consideration of non-perturbative $1/N$ corrections -- a notoriously difficult problem. Some progress~\cite{instantons,Kim} was made by recognizing that formation of the long-range spin order is equivalent to the dynamical generation of fermion mass~\cite{instantons} in $2+1$ dimensional $U(1)$ gauge theory, a problem that has been studied by particle theorists. At large-N there is a global $SU(2N)$ symmetry which combines the $N$ spin species with the two types of gapless points in the reduced Brillouin zone at $\vec{k} = (\pm \pi/2, \pi/2)$. This symmetry breaks down at finite-N as $SU(2N) \rightarrow SU(N) \otimes SU(N)$ when the fermions become gapped. Other massless excitations then arise, which can be identified as the Goldstone modes or spin-waves of the N\'eel ordered magnetic state, or equivalently as bound states of particles and holes, otherwise known as mesons. An interesting analysis of $1/N$ corrections to the $U \rightarrow \infty$ Hubbard model, with no spin exchange interaction $J$, has recently been carried out~\cite{Ziqiang}. At leading order, as expected, the system is a Fermi liquid for any doping away from half-filling. However, by working to order $1/N$ and then setting $N=2$, Fermi liquid behavior is found to be destabilized at light doping $x < 0.07$, in rough agreement with the large-N results for the $t-J$ model. \section{Phase Separation at Non-Zero Doping and Incommensurate Order} \label{sec:phase-sep} At non-zero doping there is the possibility of a staggered flux phase (SFP). In the SFP the hole occupancy is assumed to be uniform, $\langle b^\dagger_{\bf x} b_{\bf x} \rangle = N x / 2$, and the flux is reduced in magnitude from $\pi$, eventually disappearing altogether at large enough doping. Time-reversal and translational symmetries are broken in the SFP and orbital currents with real magnetic fields arise. These appear to be ruled out experimentally~\cite{2-observe}. On the other hand, phase separation into stripes had been predicted on theoretical grounds~\cite{stripe}, prior to any clear experimental observation of the phenomenon. The stability of stripes depends on the size of the hopping matrix elements beyond nearest-neighbor exchange~\cite{stripe-stability}. Phase separation solves the problem of the breaking of time-reversal and translational symmetries in the SFP. Phase separation into stripes also can explain the incommensurate, and nearly critical~\cite{Aeppli}, spin fluctuations seen in neutron scattering experiments in the 214 compound~\cite{Wells} and recently in the 123 material~\cite{Mook}. The two problems are solved simultaneously by assuming that the electron-rich region is an antiferromagnetic insulator described by the $\pi$-flux phase, with its strong tendencies towards spin ordering. The infrared divergences~\cite{instantons} at finite-N are cut-off by the limited size of the electron-rich region. Incommensurate spin peaks arise due to these finite-size effects. The tendency to phase separate can also be viewed fruitfully~\cite{Rome} in terms of a Fermi liquid close to a quantum critical point (QCP). \section{Squares, Chains, and Ladders} \label{sec:squares} Another way to study the behavior of the t-J model on the square lattice is to build up the lattice systematically from smaller subunits. For example, the four site system, a square plaquette, can be easily diagonalized. At half-filling, the spin-spin correlations in the ground state are such that the expectation value of the plaquette operator has a negative value. To be precise, in the physical $N=2$ case, the expectation values of three-spin operators such as $\langle \vec{S}_1 \cdot (\vec{S}_2 \times \vec{S_3}) \rangle$ vanish if there is no time-reversal symmetry breaking, and \begin{eqnarray} \langle \chi_{12} \chi_{23} \chi_{34} \chi_{41} \rangle &=& \frac{J^4}{16}~ \bigg{\{} \frac{1}{8} + \frac{1}{2} [ \langle \vec{S}_2 \cdot \vec{S_3} \rangle + \langle \vec{S}_2 \cdot \vec{S_4} \rangle + \langle \vec{S}_3 \cdot \vec{S_4} \rangle - \langle \vec{S}_1 \cdot \vec{S_2} \rangle - \langle \vec{S}_1 \cdot \vec{S_3} \rangle - \langle \vec{S}_1 \cdot \vec{S_4} \rangle] \nonumber \\ &-& 2~ [\langle (\vec{S}_1 \cdot \vec{S_2}) (\vec{S}_3 \cdot \vec{S_4}) \rangle + \langle (\vec{S}_1 \cdot \vec{S_4}) (\vec{S}_2 \cdot \vec{S_3}) \rangle - \langle (\vec{S}_1 \cdot \vec{S_3}) (\vec{S}_2 \cdot \vec{S_4}) \rangle] \bigg{\}} < 0\ . \end{eqnarray} Thus flux $\pi$ penetrates the plaquette. The same result holds for the $4 \times 4$ lattice with periodic boundary conditions~\cite{Runge} providing added support for the flux phase. The square lattice can also be built up chain-by-chain. A single Hubbard chain, with repulsive interactions, is a Luttinger liquid away from half-filling; at half-filling a gap develops in the charge sector, while the spin sector remains gapless in accord with the physics of a spin-1/2 quantum antiferromagnetic chain. S-wave Cooper pairing is inhibited by the strong on-site repulsion and there can be no d-wave superconducting tendencies along a single chain. Coupling two such chains with transverse hopping $t_\perp$, as shown in Fig. ~\ref{fig:ladder}, leads to more interesting possibilities. Following Fisher~\cite{LesHouches} the first step is to diagonalize the $U=0$ problem by introducing bonding ($k_y = 0$) and anti-bonding ($k_y = \pi$) orbitals: \begin{eqnarray} k_y = 0: \ \ \ \ b_{x \alpha} &\equiv& \frac{1}{\sqrt{2}}~ [c_{x \alpha 1} + c_{x \alpha 2}], \ \ \ \ \epsilon_{k_x} = -2 t_\parallel \cos(k_x) - t_\perp \nonumber \\ k_y = \pi: \ \ \ \ a_{x \alpha} &\equiv& \frac{1}{\sqrt{2}}~ [c_{x \alpha 1} - c_{x \alpha 2}], \ \ \ \ \epsilon_{k_x} = -2 t_\parallel \cos(k_x) + t_\perp\ . \end{eqnarray} \begin{figure} \epsfig{figure=ladder.eps,height=1.0in} \caption{A ladder consisting of two coupled Hubbard chains, numbered 1 and 2, with intrachain hopping in the x-direction, $t_\parallel$, and interchain hopping in the y-direction, $t_\perp$.} \label{fig:ladder} \end{figure} Now if the Hubbard repulsion is turned back on, the RG flows run away to large values suggesting that charge and spin gaps form and the problem should be studied from the strong-coupling viewpoint. Consider the limit $t_\perp \gg t_\parallel$, as it is clear that at half-filling spins on opposing sites will pair into singlets because $J_\perp \approx 4 t_\perp^2/U \gg J_\parallel \approx 4 t_\parallel^2/U$. The ground state then may be written: \begin{equation} | \Psi_0 \rangle = \prod_x \frac{1}{\sqrt{2}} [c^{\dagger \uparrow}_{x 1} c^{\dagger \downarrow}_{x 2} - c^{\dagger \downarrow}_{x 1} c^{\dagger \uparrow}_{x 2} ]~ | 0 \rangle = \prod_x \frac{1}{\sqrt{2}} [b^{\dagger \uparrow}_{x} b^{\dagger \downarrow}_{x} - a^{\dagger \uparrow}_{x} a^{\dagger \downarrow}_{x} ]~ | 0 \rangle\ . \end{equation} Thus, although there is no true off-diagonal long-range order, there is a d$_{x^2 - y^2}$ wave tendency because the last line looks like a Cooper-paired state with the sign of the pairing amplitude changing as $k_y$ switches from $0$ to $\pi$. \section{Nodal Liquid} \label{sec:nodal} A complementary approach to understanding the underdoped cuprates takes as its starting point the d-wave superconducting state~\cite{nodal,dual-vortex} with dispersion ${\cal E}(\vec{k})$ determined, using the tight-binding spectrum $\epsilon_{\vec{k}} \propto \cos(k_x) + \cos(k_y)$ and d-wave pairing gap function $ \Delta_{\vec{k}} \propto \cos(k_x) - \cos(k_y)$, by the usual BCS calculation: \begin{equation} {\cal E}(\vec{k}) = \pm \sqrt{\epsilon^2_{\vec{k}} + \Delta^2_{\vec{k}}} \propto \pm \sqrt{\cos^2(k_x) + \cos^2(k_y)}\ . \end{equation} This ``nodal liquid'' theory, as its name suggests, focuses on the gapless nodal points at $\vec{k} = (\pm \pi/2, \pm \pi/2)$. As the doping is reduced a zero-temperature quantum phase transition to a non-superconducting, but pseudo-gapped, insulating state is envisioned. The ``nodons,'' or low-energy fermion quasiparticles near the nodes, carry the same quantum numbers as the fermions in the flux phase~\cite{nodal}. Upon further reduction of the doping, antiferromagnetic order may arise~\cite{Kwon2}. \section{Isotropic RG Analysis} \label{sec:RG} The interplay between nesting, umklapp processes, and the formation of charge-density-wave (CDW), spin-density-wave (SDW), and BCS instabilities has been illustrated in a simple model of the Fermi surface which reduces the nested planes to just four Fermi points~\cite{four-points}. In a recent paper, Furukawa and Rice use a generalized version of this model to argue that the spin gap in the cuprate superconductors can be understood as a consequence of umklapp processes~\cite{Rice}. However, the whole Fermi surface can be retained. High-energy degrees of freedom are integrated out: In degenerate fermion systems this means that the inner and outer sides of the momentum shell which encloses the Fermi surface are shrunk successively~\cite{Shankar}. The RG flow can be investigated by multidimensional bosonization~\cite{Hyok-Jon,multidimensional}. The advantage of the bosonized RG calculation is that the existence of well-defined quasiparticles is not assumed {\it a priori}. Furthermore, interaction channels which possess $U(1)^\infty_k$ symmetry can be diagonalized exactly at the outset. Similar results have been obtained in the fermion basis~\cite{Virosztek,parquet,Kishine,Zanchi}. Nested Fermi surfaces support many more low-energy scattering processes than circular Fermi surfaces~\cite{multidimensional}. Some important processes are depicted in Fig. ~\ref{fig:nest}. \begin{figure} \epsfig{figure=fig2a.eps,height=2.0in} \epsfig{figure=fig2b.eps,height=2.0in} \caption{(a) A typical density-wave scattering channel which is permitted because ${\bf Q}_i + {\bf Q}_j$ is a reciprocal lattice wavevector. (b) A typical overlap between density-wave and BCS channels.} \label{fig:nest} \end{figure} Consequently, the quasiparticle lifetime $\tau \propto 1/T$ is much shorter than that in an ordinary Landau Fermi liquid ($\tau \propto 1/T^2$). Instead, the normal state is a marginal Fermi liquid and the agreement with transport data~\cite{MFL} is compelling. The tendency towards spin-density order dominates all other instabilities close to half-filling. But slightly below half-filling, because nesting is not perfect but nonetheless effective, there is competition between the SDW and BCS instabilities. As the energy cutoff $\epsilon_c$ around the Fermi surface is reduced via the RG transformation towards $\epsilon_s$, where $\epsilon_s$ is the energy deviation of the actual Fermi surface from perfect nesting, the SDW channel stops flowing but the BCS channel continues to renormalize logarithmically as its flow is independent of nesting. If the initial SDW coupling is small enough or if $\epsilon_s$ is large enough (the SDW channel does not develop an instability until $\epsilon_c \rightarrow \epsilon_s$) SDW order will not occur. Instead, d-wave superconductivity sets in at a sufficiently low temperature. \section{Consilience?} \label{sec:consilience} It should be noted that just about any theory which attempts to explain cuprate superconductivity in terms of an underlying electronic mechanism will favor d-wave order~\cite{spin-fluct}. More revealing is a comparison of other features of such theories. A summary of the four approaches discussed here, along with the more phenomenological $SO(5)$ theory of Zhang and collaborators~\cite{so5} is presented in Table \ref{tab:consilience}. Noted in the final column are theories with nodal excitations. \begin{table}[t] \caption{Comparison of different theories.\label{tab:consilience}} \vspace{0.4cm} \begin{center} \begin{tabular}{|c|c|c|l|} \hline & & & \\ Theory & Quantum Magnetism & Proximity to QCP(s) & $\vec{k} = (\pm \pi/2, \pm \pi/2)$ \\ \hline Gauge Theory & $\surd$ & $\surd$ & $\surd$ \\ Nodal Liquid & ? & $\surd$ & $\surd$ \\ RG Analysis & $\surd$ & $\surd$ & $\surd$ \\ Stripes & $\surd$ & $\surd$ & This article\\ SO(5) & $\surd$ & $\surd$ & No \\ \hline \end{tabular} \end{center} \end{table} It remains to be seen whether or not one unifying description will emerge from these pieces of the high-T$_c$ puzzle. \section*{Acknowledgments} I thank A. Houghton, H.-J. Kwon, C. Nayak and Z. Wang for recent helpful discussions. I would also like to thank the organizers of the {\it International Workshop} for their work in bringing the many foreign participants to Vietnam. This research was supported in part by the United States National Science Foundation under Grants Nos. DMR-9357613 and DMR-9712391. \section*{References}
\section{Introduction} A quantum clock is an observable capable of measuring the time duration of a physical process in a quantum system\cite{Peres,Baz}. One of the major applications discussed in this context is the measurement of the traversal time of a quantum particle through a specific region in space\cite{Baz,Buttiker,Sokolovski,Deutsch,Gasparian}: it can be performed by applying a field in this region, thereby inducing the evolution of the particle's internal states at a rate which is a known function of the field strength. If the field is weak enough, then its effect on the translational degrees of freedom of the particle is assumed to be negligible and the traversal time can then be deduced from the final internal state of the particle outside the region where the field takes effect. The first quantum clock to be discussed was based on the Larmor spin precession of an electron in a magnetic field\cite{Buttiker,Sokolovski}. By analogy, it was later proposed to use the Faraday polarization rotation of an electromagnetic wave in a birefringent crystal\cite{Deutsch} or in a magneto-refractive medium\cite{Gasparian} as a clock for photon traversal times in dielectric structures. For both Larmor precession and Faraday polarization rotation, the theoretical analysis led to the definition of a complex time variable\cite{Buttiker,Sokolovski,Gasparian}. In the case of the Larmor clock for electrons, its real and imaginary parts measure the spin rotation perpendicular and parallel to the field, respectively. In the case of the Faraday clock for photons, its real part measures the rotation of the main axis of polarization while its imaginary part measures the ellipticity of the resulting polarization. The works surveyed above have left certain open questions of fundamental importance: (i) How do the measured values of traversal times depend on the clock precision? (ii) What is the origin of the differences between traversal times measured by quantum clocks through different measuring schemes and how can they become superluminal\cite{Chiao,Spielman}? (iii) Whereas field quantization is {\it not} required for analyzing the evanescent-wave traversal (tunneling) of electromagnetic wavepackets through dielectrics or the corresponding traversal times, are there {\it nonclassical} properties of evanescent photons, which are affected by the precision of their internal quantum clock? Our purpose here is to elucidate the above questions for the Faraday rotation clock. In Section~\ref{sec:interf} we use our previously introduced explanation that evanescent-wave transmission or tunneling results from {\it destructive interference of internal traversal paths}\cite{we}, to show that {\it the mere possibility of a time-measurement by an internal clock tends to affect this interference and thus increase the transmission}. We demonstrate this fundamental result for the transmission of photons through a layered dielectric medium with Faraday rotation. In Section~\ref{sec:ttime} we discuss different measuring schemes and their relevance to the traversal time problem. We show how superluminal traversal times measured by the clock are caused by strong destructive interference of traversal paths in the evanescent-wave (tunneling) regime. In Section~\ref{sec:EPR} we discuss the effect of the Faraday clock in a dielectric structure on {\it two-photon correlations} and show that {\it correlations are not lost by the introduction of a Faraday clock}, unless the filtering effect of the clock on circular polarizations is significant. \section{Path Interference and the Faraday Clock Precision} \label{sec:interf} The Faraday effect of polarization rotation in a magneto-refractive medium is caused by the fact that the left-hand ($|+\rangle $) and right-hand ($|-\rangle $) circular polarization components with respect to an applied magnetic field have different refractive indices $n_+$ and $n_-$ in the medium. The linear polarizations $|\leftrightarrow\rangle ,|\updownarrow\rangle $ are superpositions of the two circular polarizations with equal amplitudes. While propagating a distance $x$ through the medium, the two circular polarization states $|\pm\rangle \equiv(|\leftrightarrow\rangle \pm i|\updownarrow\rangle )/\sqrt{2}$ of light at frequency $\omega$ acquire different phases $\phi_{\pm}=n_{\pm}\omega x/c$. If the transmission amplitudes of the two circular polarization states are equal, then a linearly polarized photon entering the medium will exit with another linear polarization rotated by an angle $\theta=\phi_+-\phi_-$ with respect to the initial one. However, if the transmission amplitudes are different for the two circular polarizations, then the transmitted photon is elliptically polarized. In the extreme case where only one circular-polarization component is transmitted, the medium acts as a perfect filter for circular polarizations. The use of the Faraday effect as a quantum clock is based on the fact that the phase difference between the two circular-polarization states $\phi_+ - \phi_-$ is proportional to the optical length $x\bar{n}$ in the medium, where $\bar{n}=(n_+ + n_-)/2$ is the mean refractive index. If the transmission probabilities of the two circular-polarization states are equal, then the polarization rotation of an initially linearly polarized photon is proportional to the time it spent in the medium, namely \begin{equation} \theta=\omega\frac{n_+-n_-}{\bar{n}}\frac{x}{\bar{v}} \end{equation} where $\bar{v}=c/\bar{n}$ is the mean velocity in the medium. Here we discuss the transmission of a photon through a magneto-refractive periodic dielectric structure ("a photonic band-gap structure"). The structure is made of $N$ alternating dielectric layers $1$ and $2$ with dielectric constants $n_1,n_2$ and widths $d_1,d_2$, so that the total width of the structure is $L=N(d_1+d_2)/2$. In the presence of a magnetic field each layer acquires different refractive indices for left- and right-circular polarizations. The Faraday effect can measure the traversal time through the structure if the ratio $\frac{n_+ -n_-}{\bar{n}}$ is kept constant for the two layers. The transmitted wave at $x=L$ is a superposition of many partial waves corresponding to different patterns of internal reflections between the layers. Each such pattern can be described as a path $j$, traversing $N^j_1$ times the type 1 layers and $N^j_2$ times the type 2 layers, with a certain number of internal reflections between the layers, each having an amplitude $\pm\frac{n_1-n_2}{n_1+n_2}$. We can assume that in a weak magnetic field this inter-layer reflection amplitude is approximately the same for the two circular polarizations, and therefore the total amplitude of transmission along a certain path $j$ is equal for the two polarization states. It follows from the discussion above, that the polarization state of an initially linearly polarized photon traversing a certain path $j$ in the dielectric structure remains linear and is rotated by an angle $\theta_j=\Omega(N^j_1\tau_1+N^j_2\tau_2)$, where $\tau_i=d_i \bar{n}_i/c$ ($i=1,2$) and $\Omega\equiv 2\omega\frac{n_{1+}-n_{1-}}{n_{1+}+n_{1-}}= 2\omega\frac{n_{2+}-n_{2-}}{n_{2+}+n_{2-}}$, as illustrated in Fig.~\ref{fig:illus}. The state of the transmitted photon $|\psi_{tr}(x,t)\rangle $ at $x>L$ is given by the superposition \begin{equation} |\psi_{tr}(x,t)\rangle =e^{i\omega(x-L)/c}\sum_j c_j e^{i\omega\tau_j} |\theta_j\rangle \label{psitr:exp} \end{equation} where $c_j$ is the sum over amplitudes for transmission through paths $j$, which traverse the two different dielectric layers $N^j_1$ and $N^j_2$ times and $\tau_j=N^j_1\tau_1 +N^j_2\tau_2$. The state $|\theta_j\rangle $ with $\theta_j=\Omega\tau_j$, is the linear polarization state \begin{equation} |\theta_j\rangle =\cos\Omega\tau_j|\leftrightarrow\rangle +\sin\Omega\tau_j|\updownarrow\rangle =\frac{1}{2}\left(e^{-i\Omega\tau_j}|+\rangle +e^{i\Omega\tau_j}|-\rangle \right). \end{equation} \begin{figure} \psfig{file=fclock.ps,height=8cm,width=10cm,angle=90} \caption{The Faraday clock for measuring the evanescent-wave traversal time through a magneto-refractive layered structure. The polarization rotations of the different interfering paths are proportional to the path length.} \label{fig:illus} \end{figure} A simpler form of the photon states is obtained by using the circular polarization basis. If the incident photon is in the state $|\psi_{in}\rangle =|\leftrightarrow\rangle =(|+\rangle +|-\rangle )/\sqrt{2}$, then the transmitted photon state is \begin{equation} |\psi_{tr}(x,t)\rangle =\frac{e^{i\omega(x-L)/c}}{\sqrt{2}} \left[s(\omega+\Omega)|+\rangle +s(\omega-\Omega)|-\rangle \right] \label{psixt} \end{equation} where $s(\omega)=\sum_j c_j e^{i\omega\tau_j}$ is the spectral transmission function for a monochromatic wave with frequency $\omega$. Since the two parts of the wavefunction in Eq.~(\ref{psixt}) are distinguishable by a circular polarizer, there is no interference between them and the total probability of the photon to be transmitted can be written as an {\it average} of the transmission probabilities of photons incident on a non-gyrotropic barrier at frequencies $\omega\pm\Omega$ \begin{equation} P_{tr}(\omega)=\frac{1}{2}\left[P_+(\omega) +P_-(\omega)\right] \label{Ptrw} \end{equation} where $P_{\pm}=|s(\omega \pm\Omega|^2$. This loss of interference leads to the enhancement of transmission probability in frequency bands where the transmission is low. In order to demonstrate this point, notice that if $\Omega$ is small compared to the scale of variation of $P_0(\omega)$, then we obtain from Eq.~(\ref{Ptrw}) \[ P_{\Omega}(\omega)\approx P_0(\omega)+\frac{1}{2}\Omega^2 \left. \frac{\partial^2 P_0(\nu)}{\partial \nu^2}\right|_{\nu=\omega} \] In frequency bands where the transmission is low, especially in the tunneling (evanescent) regime, the curvature $\partial^2 P_0(\nu)/\partial \nu^2$ is positive. We therefore find that the presence of the clock raises the transmission probability (as compared to the same probability in the absence of a clock), because the destructive interference of traversal paths is progressively washed out as $\Omega$ increases. \section{Time measurements by the Faraday Clock: Superluminality as Path Interference} \label{sec:ttime} What kind of measurement has to be performed in order to obtain a meaningful traversal time through the dielectric structure in Fig.~\ref{fig:illus}? If we were able to single out only transmitted photons that leave the structure in the exact state $|\theta\rangle $, then we could know what group of paths those photons went through and determine the exact traversal time through the dielectric structure. However, since the space of polarization states is spanned by only two orthogonal basis states, we are limited to the analysis of their detection probabilities rather than the determination of their state. In what follows we consider several alternative measuring schemes and discuss their relevance to the traversal time problem: A. The first alternative is to measure the number of transmitted photons $N_{orth}$ with linear polarization orthogonal to the initial polarization direction and compare it to the total number $N_{tr}$ of transmitted photons. The traversal time $\tau_{orth}$ can be defined by assuming that the polarization state of the photon after a time $\tau_{orth}$ is given by $|\theta=\Omega\tau_{orth}\rangle $. This implies that \begin{equation} N_{orth}/N_{tr}=\sin^2(\Omega\tau_{orth}). \end{equation} We then find \begin{equation} \tau_{orth}=\frac{1}{\Omega}\sin^{-1/2}\left[\frac{|s(\omega+\Omega) - s(\omega-\Omega)|^2}{4 |s(\omega)|^2}\right] \label{torth} \end{equation} In the limit $\Omega\rightarrow 0$ this becomes \begin{equation} \tau_{orth}=\left|\frac{1}{s}\frac{\partial s}{\partial \omega}\right| =\sqrt{(\partial \log |s|/\partial\omega)^2+(\partial\phi_s/\partial\omega)^2} \label{torth1} \end{equation} where $\phi_s$ is the phase of the transmission function $s=|s|e^{i\phi_s}$. The variable $\tau_{orth}$ is well known in the context of electron tunneling through potential barriers as "the B\"uttiker-Landauer time"\cite{Buttiker,Review}. B. The second alternative is to make a direct measurement of the phase difference between the two circular polarization components of the transmitted photons. This can be done by using a circular polarizer to separate the two components of the transmitted photons and then join them again by a beam-splitter as shown in Figure~\ref{fig:scheme}. If the optical length of one arm of this interferometer with respect to the other is adjustable, then the detection rate at one port of the beam-splitter as a function of the length difference between the arms satisfy the proportionality relation \begin{equation} P_{det}(\Delta x)\propto \left|s(\omega+\Omega) +is(\omega-\Omega) e^{i[\Delta\varphi+\omega\Delta x/c]}\right|^2 \end{equation} where $\Delta \varphi$ is an additional phase difference introduced by the optical components. The phase difference between the two polarization components is then given by $\phi_+ - \phi_- = \Delta\varphi+\omega \Delta x_{max}/c+\pi/2$, where $\Delta x_{max}$ is the value of $\Delta x$ which maximizes $P_{det}(\Delta x)$. The traversal time in the dielectric structure can then be extracted as \begin{equation} \tau_{ph}=\frac{\phi_+ -\phi_-}{2\Omega} \begin{array}{c} \longrightarrow \\ \Omega\rightarrow 0 \end{array} \frac{\partial \phi_s}{\partial \omega} \label{tph} \end{equation} which is the well-known value obtained in wavepacket measurements. \begin{figure} \psfig{file=scheme.ps,width=12cm,height=8cm} \caption{Measurement of the phase difference between the two circular polarization components $+$ and $-$ of a photon transmitted through a Faraday-rotation structure. } \label{fig:scheme} \end{figure} C. The third alternative is to scan the rotation angles of polarization from $\theta=0$ to $\theta=\pi$ by a linear polarizer, and define the peak rotation angle as the polarizer angle where the maximal number of counts is obtained. The counting probability for a polarizer at angle $\theta$ is given by \begin{equation} P(\theta)=\left|\langle \theta|\psi_{tr}\rangle \right|^2= \left| s_+ e^{i\theta}+s_-e^{-i\theta} \right|^2 \label{ptheta1} \end{equation} where $s_{\pm}=s(\omega\pm\Omega)$. The maximal value of this function is obtained at $\theta=(\phi_+ -\phi_-)/2$, which corresponds to the same phase-time $t_{ph}$ as given in Eq.~(\ref{tph}). In a periodically-layered dielectric structure, the evanescent-wave transmission probability $|s(\omega)|^2$ is nearly symmetric with respect to the center of the forbidden band gap (Fig.~\ref{fig:figuur}(a)) and so are the traversal times (\ref{torth1}) and (\ref{tph}) (Fig.~\ref{fig:figuur}(b)). Hence, $|s_{+}| \approx |s_{-}|$ if $\omega$ is at the band-gap center. In Ref.~\ref{we} we have shown that low transmission of a wavepacket together with abnormally short wavepacket peak traversal time is a consequence of {\it destructive interference} between consecutive partial transmitted wavepackets. In order to demonstrate the effects of interference on the clock read-out, we analyze the polarization state formed by a superposition of two linear polarization states that appear in Eq.~(\ref{psixt}). Consider the polarization state \begin{equation} |ij\rangle \equiv \lambda_i |\theta_i\rangle +\lambda_j |\theta_j\rangle , \end{equation} where $\lambda_j\equiv c_j e^{i\omega\tau_j}$ is the amplitude of the $j$th transmitted wave. As demonstrated in Figure~\ref{polfig}, if $\theta_i<\theta_j$ and $|\lambda_i|>|\lambda_j|$, it can be shown that if the interference between the terms $i$ and $j$ is predominantly destructive, i.e., when $\Re\{\lambda^*_i\lambda_j\}<0$, then the main axis of polarization of the resulting elliptic polarization state $|ij\rangle $ has an angle $\theta_{ij}<\theta_i$. In particular, if the phase difference between $\lambda_i$ and $\lambda_j$ is exactly $\pi$, such that $\lambda_j/\lambda_i=-r$, then the resulting state has linear polarization \[ \theta_{ij}=\theta_i -\arctan\left[\frac{r\sin(\theta_j-\theta_i)}{1-r\cos(\theta_j-\theta_i)} \right] \] and $\theta_{ij}<\theta_i<\theta_j$ whenever $r\cos(\theta_j-\theta_i)<1$. In general, the main axis of the elliptic polarization state of the transmitted photons is given by the maximum overlap with a linear polarization state $|\theta_m\rangle $. Using Eq.~(\ref{psitr:exp}), we can express the probability function in Eq.~(\ref{ptheta1}) as \begin{equation} P(\theta)=\left|\sum_j c_j e^{i\omega\tau_j} \cos(\theta-\theta_j)\right|^2 \end{equation} In order to obtain the value of $\theta_m$ we find the maximum of $P(\theta)$ by equating its derivative to zero. We then obtain the following angle of the main axis of polarization \[ \tan 2\theta_m=\frac{\sum_j c_j^2\sin 2\theta_j +\sum_{i\neq j}c_i c_j e^{-i\omega(\tau_i-\tau_j)}\sin(\theta_i+\theta_j)} {\sum_j c_j^2\cos 2\theta_j+\sum_{i\neq j} c_i c_j e^{-i\omega(\tau_i-\tau_j)} \cos(\theta_i+\theta_j)} \] If $\Omega$ is very small, then $\sin 2\theta\approx 2\theta$ and $\cos 2\theta\approx 1$. The clock time then reads exactly the same as the mean wavepacket traversal time\cite{we}. Hence, {\it destructive interference} between different terms $i\neq j$ gives rise to predominantly negative contributions $c_i c_j e^{-i\omega(\tau_i-\tau_j)}$, which {\it cause the effect of abnormally short or even superluminal clock read-out times}. \begin{figure} \psfig{file=tpeak.ps,height=10cm,width=10cm} \caption[]{(a) Transmission probability $|s(\omega)|^2$ for the dielectric-layered structure described in Ref.~\protect\ref{Chiao}. (b) Traversal times for the same structure: Solid line - $\tau_{ph}$ (Eq.~[\protect\ref{tph})]. Dashed - $\tau_{orth}$ (Eqs.~[\protect\ref{torth}),(\protect\ref{torth1})]. Dotted: $\tau_{amp}$ (Eq.~[\protect\ref{tamp})]. Note that the center of the band gap is approximately a symmetry point for all the plotted quantities. } \label{fig:figuur} \end{figure} \begin{figure} \psfig{file=polfig.ps,width=11cm,height=8cm} \caption{A superposition $|\theta_{12}\rangle$ (thick line) of two linear polarization states $|\theta_1\rangle , |\theta_2\rangle $ (thin lines) with different relative phases $\Delta\phi$: (a) $\Delta\phi=0$ yields $\theta_{12}>\theta_2>\theta_1$ with large amplitude; (b) $\Delta\phi=\pi$ yields $\theta_{12}<\theta_1<\theta_2$ with small amplitude; (c), (d) $\Delta\phi$ that is a non-integer fraction of $\pi$ yields elliptically polarized states.} \label{polfig} \end{figure} \section{Faraday Clock Effect in Nonlocal (EPR) Photon Correlations} \label{sec:EPR} Here we consider measurements that have never been discussed before, which pertain to non-local (Einstein-Podolsky-Rosen) correlations between entangled photons. Suppose that we repeat the Aspect experiment\cite{Aspect} using a correlated pair of photons, but insert a Faraday-rotating reflective structure (of the kind depicted in Fig.~(\ref{fig:illus})) in the way of one of the photons (Figure~\ref{fig6}). If the correlations between the transmitted photon and its twin are reduced, then the violation of Bell's inequality is expected to be weakened. To what extent does the Faraday clock affect the violation of the inequality? \begin{figure} \vspace{0.5cm} \psfig{file=epr.ps,angle=90,width=8cm,height=2cm} \vspace{0.5cm} \caption{Measurement of two-photon EPR correlations when one photon is transmitted through a Faraday-rotating reflective structure.} \label{fig6} \end{figure} Let us take the initial state of the field to be \begin{eqnarray} |\phi_{in}\rangle &=&\frac{1}{\sqrt{2}}(|\updownarrow,\updownarrow\rangle +|\leftrightarrow,\leftrightarrow\rangle )= \nonumber \\ &&=\frac{1}{\sqrt{2}}(|+,-\rangle +|-,+\rangle ) \end{eqnarray} where the first entry in the ket-vector refers to the state of the photon transmitted through the Faraday-rotating structure and travelling to detector 1 and the second entry refers to the state of the photon travelling to detector 2. Bell's inequality in the formulation of Clauser-Horne-Shimony-Holt (CHSH) makes use of the correlation function\cite{EPR} \begin{equation} E(\theta_1,\theta_2)= \frac{\langle (I^1_{\updownarrow}-I^1_{\leftrightarrow}) (I^2_{\updownarrow}-I^2_{\leftrightarrow})\rangle } {\langle (I^1_{\updownarrow}+I^1_{\leftrightarrow}) (I^2_{\updownarrow}+I^2_{\leftrightarrow})\rangle} \label{CHSH} \end{equation} where $I$ is the measured intensity, the superscripts refer to intensity measured at detectors 1 and 2 and the signs $\updownarrow,\leftrightarrow$ refer to measurements with polarizer set at angles $\theta_i$ and $\theta_i+\pi/2$ ($i=1,2$). Since detector 1 measures only the fraction of photons transmitted through the Faraday-rotating structure, the state of the photons near the detectors is given by $|\phi_{trans}\rangle =\frac{1}{\sqrt{2}} (s_+|+,-\rangle +s_-|-,+\rangle )$. We then find \begin{eqnarray} \langle I^1_{\updownarrow} I^2_{\updownarrow}\rangle &=&|\langle \phi|\theta_1,\theta_2\rangle |^2= \nonumber \\ &&=\frac{1}{2}| s_+ e^{-i\psi}+s_- e^{i\psi}|^2 \nonumber \\ \langle I^1_{\updownarrow} I^2_{\leftrightarrow}\rangle &=& \frac{1}{2}|s_+ e^{-i\psi}-s_- e^{i\psi}|^2 \end{eqnarray} where $\psi=\theta_1-\theta_2$, and similarly for the other correlations. The correlation function in Eq.~(\ref{CHSH}) becomes \begin{equation} E(\theta_1,\theta_2)=\frac{2|s_+||s_-|}{|s_+|^2+|s_-|^2} \cos(2\psi+\phi_+-\phi_-) \label{E12} \end{equation} The CHSH inequality for measurements at polarizer angles $\theta_i,\theta'_i$ is \[ |B| =|E(\theta_1,\theta_2)-E(\theta_1,\theta'_2)+E(\theta'_1,\theta'_2)- E(\theta'_1,\theta_2)| \leq 2 \] Without a Faraday rotating reflective structure we would have had $E(\theta_1,\theta_2)=\cos(2\psi)$ but with a barrier the cosine term is shifted by a certain phase and multiplied by a factor which is {\it always smaller or equal to unity} (Fig.~\ref{demons}). This factor is maximal when $|s_-|=|s_+|$, namely, when the right- and the left-circular polarizations are not separated. This happens if $\Omega$ is either very small (weak clock) or when $\omega$ {\it lies at a symmetry point} (extremum) of the $|s(\omega)|^2$ curve, e.g., if $\omega$ lies in {\it the center of the band gap} of the structure (Fig.~\ref{fig:figuur}(a)). By contrast, if the two polarizations are not equivalent, $E(\theta_1,\theta_2)$ is reduced and {\it the violation of the inequality is diminished}, indicating partial loss of the two-photon correlations. This loss of correlation is explained by the fact that the Faraday-rotating structure acts as a filter for certain polarizations and thus the transmitted photons measured at detector 1 are a sub-ensemble of the original photons, which does not represent the polarization state of the initial photon population. When $\Omega$ is small, it is easy to show from Eq.~(\ref{E12}) that the maximal value of $E(\theta_1,\theta_2)$ becomes \begin{equation} E_{max}\approx 1-2\Omega^2\tau_{amp}^2 \end{equation} where \begin{equation} \tau_{amp}=\frac{1}{|s|}\frac{\partial |s|}{\partial \omega} =\frac{\partial}{\partial\omega}\log |s| \label{tamp} \end{equation} The variable $\tau_{amp}$ measures the effectiveness of the Faraday-rotating structure as a circular polarizer. It measures the ability of the structure to separate between left- and right-circularly polarized photons. Although having units of time, it is difficult to refer to $\tau_{amp}$ any meaning connected with time duration. However, it is possible to treat $\tau_{amp}^{-1}$ as a typical frequency scale over which the transmissivity of the dielectric barrier is significantly changed. \begin{figure} \psfig{file=eprviol.ps,width=8cm,height=7cm} \caption{The maximal value of the factor B for various values of the clock precision $\Omega$ when one of the correlated photons is transmitted through a dielectric layered structure as in Fig.~\ref{fig:figuur}.} \label{demons} \end{figure} \section{Summary} We have analysed in this paper the transmission of photons through a layered dielectrtic medium with Faraday rotation and have sketched measurements of their traversal times in order to elucidate possible quantum clock mechanisms in the evanescent-wave (tunneling) regime. For the first time, both single-photon and two-photon (quantum electrodynamical) properties have been considered in this context. Destructive interference between different traversal paths has been shown to be the origin of the effect of abnormally short or even superluminal clock read-out times. Several possible schemes have been proposed for traversal-time measurements based on the Faraday-clock phase shifts. In frequency bands where the transmission is low, especially in the tunneling (evanescent-wave) regime, we have shown that the presence of the clock tends to wash out the destructive interference and thereby raise the transmission probability as compared to its counterpart in the absence of a clock. Perhaps the most intriguing subject considered here are measurements of non-local correlations between two entangled photons with a Faraday-rotation structure in the way of one of the photons. The two-photon correlation function $E(\theta_1,\theta_2)$, which is used to calculate the violation of Bell's inequality, is shifted by a phase and multiplied by a factor which is always smaller or equal to unity (Figure~\ref{demons}), as a result of the Faraday rotation. This factor is maximal when the right and left circular polarizations are not separated. This happens if the Faraday-clock rate is either very small or if the photon frequency $\omega$ lies at a symmetry point (extremum) of the $s(\omega)$ transmission curve, e.g., if $\omega$ lies at the center of a band gap. The latter result is remarkable, since it allows us to measure two-photon entanglement (Bell's inequality violation) without sacrificing the Faraday clock accuracy. By contrast, if the two polarizations are not equivalent, $E(\theta_1,\theta_2)$ is reduced and {\it the violation of the inequality is diminished}, indicating partial loss of the two-photon correlations. This loss of correlation is explained by the fact that the Faraday-rotating structure acts as a filter for certain polarizations and thus the transmitted photons measured at detector 1 are a sub-ensemble of the original photons, which does not represent the polarization state of the initial photon population. {\it $^*$ Present address: Clarendon Laboratory, Department of Physics, University of Oxford, Oxford, UK OX1 3PU} \section{Acknowledgments} This work has been supported by Minerva, ISF and EU (TMR) grants.
\section{INTRODUCTION} Narrow hypersonic jets are a ubiquitous phenomena associated with star formation. In spite of the large database of multi-wavelength observations and numerous theoretical studies, a number of fundamental questions remain unanswered about these jets. Paramount among the outstanding issues is the role of magnetic fields in the launching, collimation and propagation of Young Stellar Object (YSO) jets. The current consensus holds that the jets are formed on small scales ( $L < 10 ~AU$) via magneto-centrifugal processes associated with accretion disks (\cite{Burea96}, \cite{Shuea94}, \cite{Ouyed1997a}). At larger scales, where jet propagation rather than collimation is the issue, there is an implicit assumption that the magnetic fields involved in the launching process will remain embedded in the jets as they traverse the inter-cloud medium. Observations of narrowly collimated jets, or chains of bow-shocks, extending out to parsec scale distances (\cite{Reipurth1997}) has strengthened the viewpoint that the jet beams must ``carry their own collimators''. This is based on the possibility that without dynamically significant magnetic fields confining the beam, the jets may be disrupted by hydrodynamical instabilities (\cite{Stoneea97}, \cite{Hardeea97}). Do the magnetic fields remain embedded in the jets? That is the question we address in this paper. If we accept that jets are created via MHD forces then the only way to lose the imposed fields is via reconnection or through diffusive processes. The helical topology expected for magneto-centrifugally launched jets would not be likely to lead to large scale reconnection throughout the beam (though reconnection may have important effects on the radiative properties of the jets, \cite{Gardinea99}). Thus, if magnetic fields can be cleared out of a jet this is more likely to occur via diffusive processes. Large format CCD mosaics have recently demonstrated that YSO jets can extend over multi-parsec length scales ($L_j \approx 3$ pc). The dynamical age ($t_{dyn} = L_j/V_j$) for these jets falls in the range $t_{dyn} = 10^4 ~ - ~ 10^5$ y (\cite{Reipurth1997}, \cite{EisMun1997}). Thus, even if diffusive processes are slow, they will have a relatively long time to affect the propagation of the beams. In plasmas with less than full ionization, ambipolar diffusion is one of the most effective means of driving the rearrangement of flux relative to the plasma. Observations of optical Herbig-Haro (HH) jets have shown them to be moderately ionized, with ionization fractions ranging from $x = .5~-~.01$ (\cite{Hartiganea96}, \cite{BacEis99}), where the ionization fraction is defined as $x = n_i/n_n$ and $n_n$ and $n_i$ are the number densities of neutral and ionized particles respectively. The ionization fractions are seen to fall as the plasma propagates away from internal shocks and may, therefore, fall to even lower values in the optically invisible regions of the beam. Given the potentially low ionization fraction and the extreme ratio of jet radius to length, $R_j/L_j < 10^{-3}$, it may be possible for ambipolar diffusion to significantly alter the MHD equilibria established when the jet was launched. In this paper we wish to address the issue of ambipolar diffusion in YSO jets. Our present goal is simply to establish the characteristic time and length scales for ambipolar diffusion to operate with respect to the fundamental jet parameters. We leave its consequences to later papers. In section 2 we establish the conditions for an MHD equilibrium in a simplified model of a jet in terms of plasma, neutral and magnetic pressures. In section 3 we consider the equation for ion-neutral drift in this context and in section 4 we present an equation for the ambipolar diffusion time-scale in YSO jets. In section 5 we discuss our results in light of observations and other models. \section{Initial Configuration} We begin by considering the simplest model of a magnetized jet. We imagine a cylindrical column of material of radius $R_j$. The column is composed of ions, electrons and neutral species and is initially held together internally by a radial balance of pressure ($P_i, P_e, P_n$) and magnetic ($\vec{J} \times \vec{B})$ forces. We further assume that the column is in force balance with a (possibly magnetized) ambient medium. Thus, initially, the jet does not expand laterally ({\it i.e.,~} in the radial direction). To simplify our calculation we assume that the temperature of all species is the same ($T_i = T_e = T_n$). We further assume that the magnetic field in the jet is purely poloidal, $\vec{B} = B_z(r) \hat{e}_z$. Figure 1 shows a schematic of our model for the magnetized jet. \subsection{Global Force Balance} Our calculation is similar to derivations of the ambipolar diffusion time-scale in collapsing molecular clouds (a description of those models can be found in \cite{Spitzer78} and \cite{Mouschovias91}). Molecular clouds are cold gravitationally bound systems of extremely low ionization ($ x \approx 10^{-6} $). This implies that terms involving gas pressure and ion density can be dropped from consideration when calculating the effects of ambipolar diffusion. Gravity does not, however, play a role in maintaining the cross-sectional properties of a YSO jet and pressure forces can not be ignored. In addition, given the range of ionization fractions in these systems the ion inertia should not, {\it a priori}, be dropped from calculations. In what follows we derive a simple expression for the ambipolar diffusion time-scale (based on the ion-neutral drift velocity) including the effects of gas and plasma pressure as well as the ion density. We begin with the three fluid force balance equations for electrons ($e$), ions ($i$) and neutrals ($n$) which can be written as \begin{eqnarray} \rho_e {d \vec{v}_e \over dt} & = & - \vec{\nabla} P_e - n_e e \left( \vec{E} + {\vec{v}_e \over c} \times \vec{B} \right) + \vec{F}_{ei} + \vec{F}_{en}, \\ \rho_i {d \vec{v}_i \over dt} & = & - \vec{\nabla} P_i + n_i Z e \left( \vec{E} + {\vec{v}_i \over c} \times \vec{B} \right) + \vec{F}_{ie} + \vec{F}_{in}, \\ \rho_n {d \vec{v}_n \over dt} & = & - \vec{\nabla} P_n + \vec{F}_{ne} + \vec{F}_{ni}, \end{eqnarray} where the term $d/dt$ represents the convective derivative and $F_{12}$ represents the drag force imparted on species 1 by colliding with species 2. Note that $F_{12} = - F_{21}$. In what follows we will ignore the electron inertia term. This is valid when the time-scales of interest are long compared with the response time of the electrons. Formally this means that the dynamical time $t_{dyn}$ is longer than either the electron cyclotron period $2\pi/\Omega_{ec}$ or the electron plasma period $2\pi/\omega_{ep}$ (\cite{elliot93}) as is the case in YSO jets with fields greater than the $\mu G$ level. We will also neglect the electron collisional coupling to the the neutrals, $F_{en}$, given the difference in masses, $M_e << M_i,M_n$ (see footnote \#4 \cite{Mous1996}, Ciolek private communication 1999). With these caveats the addition of equations 1 and 2 yields the following expression: \begin{equation} \rho_i {d \vec{v}_i \over dt}= - \vec{\nabla} (P_i + P_e) + (n_i Z e - n_e e) \vec{E} + \left( {n_i Z e \vec{v}_i - n_e e \vec{v}_e \over c} \times \vec{B} \right) + \vec{F}_{in}. \label{add12} \end{equation} \noindent Assuming charge neutrality ($n_i Z = n_e$) yields the following form for the current density, \begin{equation} \vec{J} = n_i Z e \vec{v}_i - n_e e \vec{v}_e = n_e e (\vec{v}_i - \vec{v}_e), \label{currdef} \end{equation} \noindent and the term with the electric field drops out of equation \ref{add12}. Making use of Amp\'ere's law, the Lorentz force can be decomposed into two terms, \begin{equation} {1 \over c}(\vec{J} \times \vec{B}) = - \vec{\nabla}_\perp {B^2 \over 8 \pi} + {B^2 \over 4 \pi R_c} \hat{n} \label{lforce} \end{equation} \noindent where $\hat{n}$ is a unit vector directed toward the local center of curvature of the field line, $R_c$ is the local radius of curvature, and $\vec{\nabla}_{\perp}$ is the gradient normal to field lines. Our choice of a $\vec{B} = B_z(r) \hat{e}_z$ implies that the second term in equation \ref{lforce} is zero. We are left, therefore, with only the first term which can be identified as the gradient of magnetic pressure, $P_B = B^2/8\pi$. Finally, if we add the momentum equations for all three species, again ignoring the electron inertia we arrive at the following equation, \begin{equation} \rho_i {d \vec{v}_i \over dt} + \rho_n {d \vec{v}_n \over dt} = - \vec{\nabla} (P_i + P_e + P_n + P_B). \label{forcebal} \end{equation} \noindent In what follows we need only consider the radial component of the force equation. Thus if, \begin{equation} {\partial \over \partial r} (P_i + P_e + P_n + P_B) = 0, \label{Gpbal} \end{equation} \noindent and ${v}_{i,r} = {v}_{n,r} = {v}_{e,r} = 0$ (where the $v_r$'s refer to radial velocities) at $t = 0$ then the entire configuration begins in equilibrium. It is, however, what we shall call a {\it quasi-equilibrium} since each species will be accelerated separately by the terms on the RHS of equations 1 - 3. The plasma and the neutrals will attempt to move past each other as each responds to its own forces. It is only the collisional drag $F_{ij}$ between species which holds the initial configuration together. This is exactly the situation which occurs in a molecular cloud where the cloud can be supported for some period of time by its magnetic field. As in the molecular cloud case, the quasi-equilibrium in the jet can not persist indefinitely and the initial configuration will change on a time-scale $t_{ad} = R_j/v_d$ where $v_d$ is the ion-neutral drift velocity $v_d = v_n - v_i$. We seek to calculate this time-scale. Note that equation \ref{Gpbal} implies that the pressure distributions of the ions and neutrals is not independent. \begin{equation} {\partial \over \partial r} P_n = - {\partial \over \partial r} (P_i + P_e + P_B) \label{pndep} \end{equation} \noindent We will use this relation below in our calculation of $t_{ad}$. \section{The Ion-Neutral Drift} If we divide equation \ref{add12} by $\rho_i$ and subtract from it equation 3 (itself divided by $\rho_n$) we find, \begin{equation} {d \vec{v}_i \over dt} - {d \vec{v}_n \over dt} = - {1 \over \rho_i} \vec{\nabla} (P_i + P_e + P_B) + {1 \over \rho_n} \vec{\nabla} P_n + \vec{F}_{in} \left({\rho_n + \rho_i \over \rho_n \rho_i}\right). \label{that} \end{equation} \noindent If we now assume that the collisions will quickly bring the ion-neutral drift to a steady velocity, {\it i.e.,~} the relative accelerations approach zero after a few collisions, we have the following radial balance equation between collisions and hydro-magnetic forces, \begin{equation} F_{in}\left({\rho_n + \rho_i \over \rho_n \rho_i}\right) = {1 \over \rho_i} {\partial \over \partial r} (P_i + P_e + P_B) - {1 \over \rho_n} {\partial \over \partial r} P_n. \label{driftbal} \end{equation} \noindent Substituting equation \ref{pndep} into the above expression we find the densities cancel, leading to \begin{equation} {F}_{in} = {\partial \over \partial r} (P_i + P_e + P_B). \label{fcon1} \end{equation} \noindent We now make the approximation that the scale of the gradients in the jet are such that \begin{equation} {\partial P_k \over \partial r} \approx {P_{k} \over R_j} \label{pscale} \end{equation} \noindent where $P_k$ is is the characteristic scale for that pressure of the k-th component of the configuration. We also use the definition of the plasma $\beta$ parameter, $\beta = P_g/P_B$, where $P_g = P_i + P_e$. Thus \begin{eqnarray} F_{in} & = & {P_{B} \over R_j} (\beta +1) \\ & = & { {B}^2 \over 8 \pi R_j} (\beta +1). \end{eqnarray} \noindent To reduce this equation further we must consider the form of the collisional force. $F_{in}$ can be written as, \begin{equation} \vec{F}_{in} = \rho_i \rho_n {\langle\sigma w \rangle_{in} \over M_i + M_n} (\vec{v_n} - \vec{v_i}), \label{fdef} \end{equation} \noindent where $\langle\sigma w \rangle_{in} \approx 3 \times 10^{-9} ~cm^3 ~s^{-1}$ is the average collision rate (\cite{Spitzer78}). We can now solve for the ion-neutral drift velocity. \begin{equation} v_d = {M_i + M_n \over 8 \pi} {1 \over \langle\sigma w \rangle_{in} } \left( {B^2 \over \rho_i \rho_n R_j} \right) (\beta + 1). \label{vdeq} \end{equation} \noindent Thus the time-scale for changes in the initial configuration can be written as \begin{equation} t_{ad} = {8 \pi \langle\sigma w \rangle_{in} M_n M_i \over M_i + M_n} x \left({n _n R_j \over B} \right)^2 ( \beta + 1)^{-1}, \label{vdeq1} \end{equation} \noindent where we have used $x = n_i/n_n$. The above expression gives the ambipolar diffusion time-scale in terms of the fundamental parameters in the jet ($x, n_n, R_j$ and $B$). \section{Ambipolar Diffusion in YSO Jets} Of the four variables in equation \ref{vdeq1} only first three ($x$, $n_n$, $R_j$) are well characterized for HH jets. From observations typical values are: $.01 \le x \le .1$; $10^3 \le n_n/~cm^{-3} \le 10^4$; $1 \times 10^{15}~ \le R_j/~cm \le 5 \times 10^{15}$ (\cite{BacEis99}). Magnetic fields however are not so easily categorized. To date there have been only a handful of measurements of fields in YSO jets (\cite{Ray1997}). Thus, it would be better to cast equation \ref{vdeq1} in terms of a parameter such as the temperature which has been measured in many YSO jets. Typical values in optical jets are $5 \times 10^3 ~K < T < 3 \times 10^4 ~K$, (\cite{BacEis99}). To express equation \ref{vdeq1} in terms of T we use our assumption that the electron and ion temperature is the same and, furthermore, assume that the plasma is mainly composed of Hydrogen atoms. Thus \begin{eqnarray} M_i & = & M_n \\ n_i & = & n_e \\ P_e = n_e k T & = & n_i k T = P_i \end{eqnarray} \noindent The magnetic field can now be expressed as \begin{equation} B^2 = 8\pi \frac{P_e + P_i}{\beta} = 16 \pi n_i k T {1 \over \beta}, \end{equation} \noindent and the ambipolar diffusion time-scale becomes \begin{equation} t_{ad} = {M_n \langle\sigma w \rangle_{in} \over 4 k } \left( { n_n R_j^2 \over T_j} \right) \left( {\beta \over \beta + 1} \right). \label{vdeq2} \end{equation} \noindent Upon normalization the ambipolar diffusion time-scale can be written as \begin{eqnarray} t_{ad} & = & 28,904 \left( { n_n \over 10^3 cm^{-3}} \right) \left( { R_j \over 10^{15} cm} \right)^2 \left( { 10^4 K \over T_j} \right) Q(\beta) ~ y \\ Q(\beta) & = & \left( {\beta \over \beta + 1} \right) \label{vdeq3} \end{eqnarray} \noindent Note that $Q(\beta)$ is positive with an asymptotic value of $1$ as $\beta \rightarrow \infty$. This shows that the ion-neutral drift can be driven by gas pressure forces alone if the individual species' pressure gradients have different signs, an artificial situation not likely to be encountered in real jets. Detailed calculations of equilibrium magnetically confined jets arising from accretion disks indicate that $\beta$ can become quite low in some regions of the jet. As an example, we provide in Figure 2 a plot of $\beta$ versus radius for an equilibrium MHD jet. This model was calculated via the Given Inner Geometry method of Lery {\it et al.}~ (\cite{Leryea98}, \cite{Leryea99}) for determining the asymptotic structure of magneto-centrifugally launched flows. Figure 2 shows that $\beta$ can drop to values below $\beta = 1$ in the {\it inner} regions of the jet ($r < R_j$). In Figure 3 and 4 we show surface plots of $t_{ad} ~vs~ \beta$ for $\beta = 1$ and $\beta = .1$. From equations 25 and 26 and these figures it is clear that depending on the conditions in the jets, $t_{ad}$ can become as large as $10^6$ y and as small as $10^3$ y. For jet parameters in the middle of the expected range of variation we find $t_{ad}$ of order $10^5$ to $10^4$ y. What is noteworthy about these results is a significant region of parameter space exists where the dynamical time-scale for YSO jets is of order of, or greater than, the ambipolar diffusion time: $t_{dyn} > t_{ad}$. {\it Thus ambipolar diffusion is likely to play a role in the dynamics of large scale YSO jets and outflows}. We discuss the consequences of this conclusion in the next section. \section{Discussion and Conclusions} Our conclusion that ambipolar diffusion time-scales can be comparable to jet dynamical lifetimes raises a number of intriguing issues. The first is the most obvious. Ambipolar diffusion will rearrange the mass to flux ratio in the jet and alter any initial equilibrium between the plasma, neutrals and magnetic field. Our analysis is too simplified to yield conclusions about how the jet will evolve in the presence of ambipolar diffusion. In our analysis we did not consider the effect of toroidal fields. The hoop stresses associated with toroidal fields will the pull ions towards the jet axis. Thus depending on the orientation of the magnetic pressure gradients, magnetic forces (pressure and tension) can either compete or apply forces in the same direction. Including a toroidal component does not, in general, change the order of magnitude of the ambipolar diffusion time-scale, however, it can change the direction in which the ions are pushed. The plasma and magnetic field may bleed out of the jet leaving only the neutrals or, conversely, strong hoop stresses could draw the plasma and field in towards the jet axis. In both cases, however, the neutrals will be free to contract or expand depending on their pressure distribution relative to the ambient medium. Thus one expects a potentially significant rearrangement of the jet's cross sectional properties when ambipolar and jet time-scales are comparable. The consequences of ambipolar diffusion on long term jet propagation should, therefore, be investigated in detail. The possibility that many jets will have $t_{dyn} \approx t_{ad}$ is suggestive for the general issue of YSO outflows. There remains considerable debate over the connection between HH jets and molecular outflows (\cite{MasCher93}, \cite{Cabrit97}). The discovery of parsec scale ``superjets'' (\cite{Reipurth1997}, \cite{EisMun1997}) has strengthened the case for jets as the source of molecular outflows by extending jet lifetimes. Still, most outflows have larger dynamical timescales than even the superjets by a factor of at least a few. One may wonder then why the jets have shorter lifetimes then the molecular outflows. Our results suggest that in some systems at least it is ambipolar diffusion which sets the lifetime of the visible collimated jet. Converting to consideration of length scales, if one takes the minimum ambipolar diffusion time from our calculation, $t_{ad}\approx 10^3 ~y$ and the minimum characteristic jet velocity, $V_j \approx 10^2 ~km/s$ one finds the minimum distance where ambipolar diffusion becomes effective: $D_{ad} \approx 0.3 ~pc$ Our results indicate that ambipolar diffusion should not be effective before a jet reaches this distance. Moreover for typical values $t_{ad} \approx 10^4 ~y$ and $V_j \approx 3 \times 10^2 ~km/s$ one finds $D_{ad} \approx 3 ~pc$. It is indeed noteworthy that this value corresponds well with the distance where the longest jets start to fade away. When $t_{dyn}$ becomes comparable to $t_{ad}$ the beam may no longer be confined by magnetic forces. \acknowledgements We wish to thank Glenn Ciolek, Larry Helfer and Al Simon for their help with this project. We also acknowledge the helpful comments of Tom Ray, Francesca Bacciotti and Alex Raga. This work was supported by NSF Grant AST-0978765 \begin{figure} \epsfxsize=2.0in \centerline{\epsfbox{Fig1.eps}} \figcaption{Cartoon of Jet Model. Shown is a section of a magnetized jet composed of a plasma, neutral gas and magnetic field oriented along the axis of the jet. The pressure and field distributions are a function of cylindrical radius only.} \end{figure} \begin{figure} \epsfxsize=6.0in \centerline{\epsfbox{Fig1.eps}} \figcaption{Plasma beta vs jet radius for MHD equilibrium jet. Jet model taken from calculations of Lery {\it et al.}~ 1999. Note that lowest values of beta occur within} \end{figure} \begin{figure} \epsfxsize=6.0in \centerline{\epsfbox{Fig3.eps}} \figcaption{The Ambipolar diffusion time-scale as a function of temperature (in units of 10,000 K) and density (in units of 1000 $cm^{-3}$) when $\beta = 1$. The jet radius is $2 \times 10^{15}$ cm.} \end{figure} \begin{figure} \epsfxsize=6.0in \centerline{\epsfbox{Fig4.eps}} \figcaption{The Ambipolar diffusion time-scale as a function of temperature (in units of 10,000 K) and density (in units of 1000 $cm^{-3}$) when $\beta = .1$. The jet radius is $2 \times 10^{15}$ cm.} \end{figure}
\section{Introduction} For any Lagrangian based field theory a 'metrical' energy-momentum tensor is defined by \begin{equation} T^{\mu\nu} = -\frac 2{\sqrt{-g}} \frac{\delta L}{\delta g_{\mu\nu}}, \label{1} \end{equation} where $g_{\mu\nu}$ is the metric tensor, $L$ is the Lagrangian density of the field and $g$ is the determinant of $g_{\mu\nu}$. The field equations (the Euler-Lagrange equations) are derived by applying the variational principle: \begin{equation} \frac {\delta L}{\delta \phi_A}=0, \end{equation} where $\phi_A$ are field variables. In the geometrical description of GR, one identifies the gravitational field with the geometry of the curved space-time, and the metric tensor $g_{\mu\nu}$ plays the dual role. On one side, the metric tensor $g_{\mu\nu}$ defines the metrical relations in the space-time, on the other side, the components of the $g_{\mu\nu}$ are regarded as the gravitational field variables. The variation of $L$ with respect to $g_{\mu\nu}$ gives the Einstein equations. If one defines the gravitational energy-momentum tensor according to (\ref{1}), it becomes equal to zero due to the Einstein field equations. The quantities which are normally being used are various energy-momentum pseudotensors. They are known to be unsatisfactory. The lack of a rigorously defined energy-momentum tensor for the gravitational field becomes especially acute in problems such as quantization of cosmological perturbations. We believe that the difficulty lies in the way we treat the gravitational field, not in the nature of gravity as such. The Einstein gravity can be perfectly well formulated as a field theory in the Minkowski space-time. This allows us to introduce a perfectly acceptable energy-momentum tensor for the gravitational field. \section{The field theoretical approach to gravity} The field theoretical approach treats gravity as a nonlinear tensor field $h^{\mu\nu}$ given in the Minkowski space-time. This approach has a long and fruitful history (see for example \cite{Feynman,Deser,GPP}). However, the previous work was focused on demonstrating that the Einstein equations are, in a sense, most natural and unavoidable, rather than on any practical applications of this approach. We will follow a concrete representation developed in \cite{GPP}. The gravitational field is denoted by $h^{\mu\nu}$, and the metric tensor of the Minkowski space-time by $\gamma^{\mu\nu}$ . The Christoffel symbols associated with $\gamma^{\mu\nu}$ are denoted by $C^{\tau}_{\ \mu\nu}$, the covariant derivatives with respect to $\gamma^{\mu\nu}$ by ";" and the curvature tensor of the Minkowski space-time is \begin{equation} \breve R_{\alpha\beta\mu\nu}(\gamma^{\rho\sigma})=0. \label{r} \end{equation} \subsection{The variational principle} The action for the gravitational field is \begin{equation} S^g=- \frac1{2c\kappa} \int L^g \;d^4x, \nonumber \end{equation} where $\kappa = 8\pi G / c^4$. The energy-momentum tensor for the gravitational field is defined in the traditional manner, as the variational derivative with respect to the metric tensor $\gamma_{\mu\nu}$: \begin{equation} \kappa t^{\mu\nu}|_v = -\frac 1{\sqrt{-\gamma}} \frac{\delta L^g}{\delta \gamma _{\mu\nu}}, \label{4} \end{equation} The field equations are derived by applying the variational principle with respect to the field variables $h^{\mu\nu}$ (gravitational potentials). It is convenient (but not necessarily) to consider the generalised momenta $P^{\alpha}_{\ \mu\nu}$, canonically conjugated to the generalised coordinates $h^{\mu\nu}$, as independent variables . This is an element of the Hamiltonian formalism, which is known also as the first order variational formalism. So, the field equations, in the framework of the first order formalism, are \begin{equation} \frac{\delta L^g}{\delta h^{\mu\nu}}= 0, \label{5} \end{equation} \begin{equation} \frac{\delta L^g}{\delta P^{\tau}_{\ \mu\nu}} =0. \label{6} \end{equation} The concrete Lagrangian density for the gravitational field used in \cite{GPP} is given by: \begin{equation} L^{g}= \sqrt{-\gamma} \left({h^{\rho\sigma}}_{;\alpha}P^{\alpha}_{\ \rho\sigma} - \frac 1{2}{\Omega^{\rho\sigma\alpha\beta}}_{\omega\tau}P^{\tau}_{\ \rho\sigma}P^{\omega}_{\ \alpha\beta} \right),\label{7} \end{equation} where \begin{eqnarray} {\Omega^{\rho\sigma\alpha\beta}}_{\omega\tau} &\equiv& \frac 1{2} [ (\gamma^{\rho\alpha}+h^{\rho\sigma}){Y^{\sigma\beta}}_{\omega\tau} + (\gamma^{\sigma\alpha}+h^{\sigma\alpha}){Y^{\rho\beta}}_{\omega\tau}+ \nonumber\\ & &(\gamma^{\rho\beta}+ h^{\rho\beta}){Y^{\sigma\alpha}}_{\omega\tau}+ (\gamma^{\rho\alpha} + h^{\rho\alpha}){Y^{\sigma\beta}}_{\omega\tau}]\label{o} \end{eqnarray} and $$ {Y^{\rho\alpha}}_{\sigma\beta} \equiv \delta^{\rho}_{\sigma}\delta^{\alpha}_{\beta} -\frac 1{3} \delta^{\rho}_{\beta} \delta^{\alpha}_{\sigma}. $$ \subsection{The field equations} By direct calculation of the variational derivatives of the Lagrangian density (\ref{7}) one can obtain the field equations: \begin{equation} \frac1{\sqrt{-\gamma}}\frac{\delta L^g}{\delta h^{\mu\nu}}\equiv - \left( P^{\alpha}_{\ \mu\nu ;\alpha} + P^{\alpha}_{\ \mu\beta} P^{\beta}_{\ \nu\alpha} -\frac1{3} P_{\mu} P_{\nu}\right)= 0, \label{8} \end{equation} \begin{equation} \frac1{\sqrt{-\gamma}}\frac{\delta L^g}{\delta P^{\tau}_{\ \mu\nu}}\equiv {h^{\mu\nu}}_{;\tau} - { \Omega^{\mu\nu\alpha\beta}}_{\omega\tau}P^{\omega}_{\ \alpha\beta}=0,\label{9} \end{equation} where $P_{\rho}\equiv P^{\alpha}_{\ \rho\alpha}$. Equation (\ref{9}) provides the link between $P^{\alpha}_{\ \mu\nu}$ and $h^{\mu\nu}$: \begin{equation} {h^{\mu\nu}}_{;\tau} = {\Omega^{\alpha\beta\mu\nu}}_{\tau\omega}P^{\omega}_{\ \alpha\beta},\label{9b} \end{equation} One can also resolve equation (\ref{9b}) in terms of $P^{\tau}_{\ \mu\nu}$: \begin{equation} P^{\tau}_{\ \mu\nu} = {\Omega^{-1}_{\rho\sigma\mu\nu}}^{\tau\omega} {h^{\rho\sigma}}_{;\omega}, \label{p} \end{equation} where ${\Omega^{-1}_{\rho\sigma\mu\nu}}^{\tau\omega}$ is the inverse matrix to the matrix ${\Omega^{\alpha\beta\mu\nu}}_{\tau\omega}$, namely \begin{equation} {\Omega^{\mu\nu\alpha\beta}}_{\omega\tau} {\Omega^{-1}_{\rho\sigma\mu\nu}}^{\tau\psi} \equiv \frac 1{2} \delta^{\psi}_{\omega}(\delta^{\alpha}_{\rho}\delta^{\beta}_{\sigma}+ \delta^{\alpha}_{\sigma}\delta^{\beta}_{\rho}).\label{o-1} \end{equation} The field equations (\ref{8}) could be also derived using the Lagrangian formalism, known also as a second order variational formalism. To implement this one has to consider $P^{\alpha}_{\ \mu\nu}$ as known functions of $h^{\mu\nu}$ and $h^{\mu\nu}\ _{;\alpha}$ (see (\ref{p})) and substitute $P^{\tau}_{\ \mu\nu}$ into the Lagrangian (\ref{7}). After performing this substitution, the Lagrangian takes the following form \begin{equation} L^{g}=\frac 1 {2} \sqrt{-\gamma} {\Omega^{-1}_{\rho\sigma\alpha\beta}}^{\omega\tau} {h^{\rho\sigma}}_{;\tau} {h^{\alpha\beta}}_{;\omega} , \label{3n} \end{equation} which is explicitely quadratic in term of "velocities" ${h^{\mu\nu}}_{;\tau}$. The field equations, in framework of the second order variational formalism, are \begin{equation} \frac{\delta L^g}{\delta h^{\mu\nu}}= 0 \label{5n} \end{equation} These equations are exactly the same as equations (\ref{8}), if one takes into account the link (\ref{p}). \subsection{Connection to the geometrical GR} Equations (\ref{8}), (\ref{9}) are fully equivalent to the Einstein equations in the geometrical approach. To demonstrate the equivalence, one has to introduce new quantities $g^{\mu\nu}$ according to the definition: \begin{equation} \sqrt{-g}g^{\mu\nu}=\sqrt{-\gamma}(\gamma^{\mu\nu}+ h^{\mu\nu})\label{g} \end{equation} and interpret $g_{\mu\nu}$ as the metric tensor of a curved space-time. The matrix $g_{\mu\nu}$ is the inverse matrix to $g^{\mu\nu}$ \begin{equation} g_{\mu\alpha}g^{\nu\alpha}=\delta^{\nu}_{\mu} \label{g-1} \end{equation} and $g$ is the determinant of $g_{\mu\nu}$. The Christoffel symbols $\Gamma^{\alpha}_{\ \mu\nu}$ constructed from $g_{\mu\nu}$, with (\ref{p}) and (\ref{g}) taken into account, have the following form \begin{equation} \Gamma^{\alpha}_{\ \mu\nu} = C^{\alpha}_{\ \mu\nu} -P^{\alpha}_{\ \mu\nu} + \frac 1{3} \delta^{\alpha}_{\mu} P_{\nu} + \frac 1{3} \delta^{\alpha}_{\nu} P_{\mu}. \label{G} \end{equation} The vacuum Einstein's equations \begin{equation} R_{\mu\nu} = {\Gamma^{\alpha}_{\ \mu\nu}}_{,\alpha} -\frac1{2} \Gamma_{\mu,\nu} -\frac1{2} \Gamma_{\nu,\mu} + \Gamma^{\alpha}_{\ \mu\nu}\Gamma_{\alpha}- \Gamma^{\alpha}_{\ \mu\beta}\Gamma^{\beta}_{\ \nu\alpha}=0, \label{ein} \end{equation} with $\Gamma^{\alpha}_{\ \mu\nu}$ taken from (\ref{G}), reduce to \begin{equation} R_{\mu\nu} = \breve R_{\mu\nu}- \left( P^{\alpha}_{\ \mu\nu ;\alpha} + P^{\alpha}_{\ \mu\beta} P^{\beta}_{\ \nu\alpha} -\frac1{3} P_{\mu} P_{\nu}\right)= 0, \end{equation} which are exactly the field equations (\ref{8}), because $\breve R_{\mu\nu}=0$. \section{The energy-momentum tensor for the gravitational field} In general, the energy-momentum tensor derived from the Lagrangian (\ref{7}) according to the definition (\ref{4}) contains the second order derivatives of the gravitational potentials $h^{\mu\nu}$: \begin{eqnarray} \kappa t^{\mu\nu}|_v &=& \frac1{2} \gamma^{\mu\nu}{h^{\rho\sigma}}_{;\alpha} P^{\alpha}_{\ \rho\sigma} + [\gamma^{\mu\rho}\gamma^{\nu\sigma} - \frac1{2}\gamma^{\mu\nu}(\gamma^{\rho\sigma} + h^{\rho\sigma})] (P^{\alpha}_{\ \rho\beta}P^{\beta}_{\ \sigma\alpha} - \nonumber \\ & & \frac1{3} P_{\rho}P_{\sigma}) + Q^{\mu\nu}, \label{tv} \end{eqnarray} where \begin{equation} Q^{\mu\nu}= \frac1{2}(\delta^{\mu}_{\rho}\delta^{\nu}_{\sigma}+ \delta^{\nu}_{\rho}\delta^{\mu}_{\sigma}) [-\gamma^{\rho\alpha} h^{\beta\sigma}P^{\tau}_{\ \alpha\beta} +(\gamma^{\alpha\tau}h^{\beta\rho} - \gamma^{\alpha\rho}h^{\beta\tau})P^{\sigma}_{\ \alpha\beta}]_{;\tau}. \end{equation} Some of the second order derivatives (but not all) can be excluded by using the field equations. We regard an energy-momentum tensor physically satisfactory if it does not depend on the second derivatives of the field potentials. The remaining freedom in the Lagrangian (\ref{7}), which preservs the field equations (\ref{8}), (\ref{9}), allows us to build such an object. Our aim is to construct a symmetric conserved energy-momentum tensor which does not contain the higher than the first order derivatives of the gravitational potentials $h^{\mu\nu}$. These are the demands which should be met by a fully acceptable energy-momentum tensor. To derive such an energy-momentum tensor we use the most general Lagrangian density consistent with the Einstein equations (\ref{8}), (\ref{9}) and with the variational procedure: \begin{eqnarray} L^{g} = \sqrt{-\gamma} \left[{h^{\rho\sigma}}_{;\alpha}P^{\alpha}_{\ \rho\sigma} - \frac 1{2}{\Omega^{\rho\sigma\alpha\beta}}_{\omega\tau}P^{\tau}_{\ \rho\sigma}P^{\omega}_{\ \alpha\beta}\right] + \sqrt{-\gamma}\Lambda^{\alpha\beta\rho\sigma} \breve R_{\alpha\rho\beta\sigma},\label{l'} \end{eqnarray} where $\breve R_{\alpha\rho\beta\sigma}$ is the curvature tensor (\ref{r}) constructed from $\gamma_{\mu\nu}$. We have added zero to the original Lagrangian, but this is a typical way of incorporating a constraint (in our case, $\breve R_{\alpha\rho\beta\sigma}=0$) by means of the undetermined Lagrange multipliers. The multipliers $\Lambda^{\alpha\mu\beta\nu}$ form a tensor which depends on $\gamma^{\mu\nu}$ and $h^{\mu\nu}$. The added term affects the energy-momentum tensor, but does not change the field equations. The field equations are derived from (\ref{l'}) using the variational principle as it was described above (see (\ref{5}), (\ref{6})). The equations derived from the new Lagrangian are exactly the same as original equations (\ref{8}), (\ref{9}). The energy-momentum tensor directly derived from (\ref{l'}) according to (\ref{4}) is now modified as compared with (\ref{tv}): \begin{eqnarray} \kappa t^{\mu\nu}|_v &=& \frac1{2} \gamma^{\mu\nu}{h^{\rho\sigma}}_{;\alpha} P^{\alpha}_{\ \rho\sigma} + [\gamma^{\mu\rho}\gamma^{\nu\sigma} - \frac1{2}\gamma^{\mu\nu}(\gamma^{\rho\sigma} + h^{\rho\sigma})] (P^{\alpha}_{\ \rho\beta}P^{\beta}_{\ \sigma\alpha} - \nonumber \\ & & \frac1{3} P_{\rho}P_{\sigma}) + Q^{\mu\nu} + \frac1{\sqrt{-\gamma}} \frac{\partial [\sqrt{-\gamma} \gamma_{\alpha\tau} \Lambda^{\tau\beta\rho\sigma}]}{\partial \gamma_{\mu\nu}} \breve R^{\alpha}_{\ \rho\beta\sigma} - \nonumber\\ & & (\Lambda^{\nu\beta\mu\alpha} + \Lambda^{\mu\beta\nu\alpha})_{;\alpha ;\beta}.\label{tv'} \end{eqnarray} This tensor still contains the second derivatives of the gravitational potentials $h^{\mu\nu}$, but the multipliers $\Lambda^{\alpha\beta\rho\sigma}$ can be chosen in such a way that the remaining terms which contain the second derivatives of $h^{\mu\nu}$ (and which could not be excluded by using the field equations) can now be removed. We have shown that the unique choice of $\Lambda^{\alpha\beta\rho\sigma}$ is \begin{equation} \Lambda^{\alpha\beta\rho\sigma}= -\frac 1{4} (h^{\alpha\beta}h^{\rho\sigma} -h^{\alpha\sigma}h^{\beta\rho}). \end{equation} As a result we have obtained the energy-momentum tensor (which we call the true energy-momentum tensor), which is \begin{itemize} \item 1) symmetric, \item 2) conserved, ${t^{\mu\nu}}_{;\nu}=0$, as soon as the field equations are satisfied, \item 3) does not contain any derivatives of $h^{\mu\nu}$ higher than the first order. \end{itemize} It is necessarily to emphasize that $t^{\mu\nu}$ is a tensor (not a pseudotensor), so that it transforms as a tensor under arbitrary coordinate transformations. The true energy-momentum tensor has the following final form: \begin{eqnarray} \kappa t^{\mu\nu} &=& \frac1{4} [2 {h^{\mu\nu}}_{;\rho} {h^{\rho\sigma}}_{;\sigma} -2{h^{\mu\alpha}}_{;\alpha}{h^{\nu\beta}}_{;\beta} + 2g^{\rho\sigma}g_{\alpha\beta} {h^{\nu\beta}}_{;\sigma} {h^{\mu\alpha}}_{;\rho} + \nonumber \\ & & g^{\mu\nu}g_{\alpha\rho} {h^{\alpha\beta}}_{;\sigma} {h^{\rho\sigma}}_{;\beta}- 2g^{\mu\alpha}g_{\beta\rho} {h^{\nu\beta}}_{;\sigma}{h^{\rho\sigma}}_ {;\alpha} - \nonumber \\ & & 2g^{\nu\alpha} g_{\beta\rho} {h^{\mu\beta}}_{;\sigma} {h^{\rho\sigma}}_{;\alpha} + \frac1{4}(2g^{\mu\delta}g^{\nu\omega} -g^{\mu\nu}g^{\omega\delta}) (2g_{\rho\alpha}g_{\sigma\beta} - \nonumber \\ & & g_{\alpha\beta}g_{\rho\sigma}) {h^{\rho\sigma}}_{;\delta} {h^{\alpha\beta}}_{;\omega}],\label{49} \end{eqnarray} where $g_{\alpha\beta}$ and $g^{\alpha\beta}$ are defined by (\ref{g}) and (\ref{g-1}). It is important to point out that $t^{\mu\nu}$ can be numerically related with the Landau-Lifshitz pseudotensor \cite{LL}. If one introduces the quantities $g^{\mu\nu}$ according to (\ref{g}) and uses the Lorentzian coordinate frame of reference ($\gamma_{00}=1, \gamma_{11}=\gamma_{22}=\gamma_{33}=-1$) one obtains the relationship between $t^{\mu\nu}_{LL}$ (the Landau-Lifshitz pseudotensor) and $t^{\mu\nu}$: \begin{equation} t^{\mu\nu} = (-g) t^{\mu\nu}_{LL}. \end{equation} \section{Gravitational field with matter sources} So far we were working with gravitational equations without matter sources, but all procedure cosidered above is also true for gravitational field in presence of matter. In order to maintain equivalence with the Einstein equations in its geometrical form one needs to ensure the universal coupling of gravity to other physical fields (matter sources). The universal coupling between gravity and matter is presented as the following functional dependence of the matter Lagrangian density: \begin{equation} L^m=L^m(\sqrt{-\gamma} (\gamma^{\mu\nu} + h^{\mu\nu}); \phi_A;\phi_{A,\alpha}), \label{57} \end{equation} where $\phi_A$ is an arbitrary matter field. Then, the energy-momentum tensor $\tau^{\mu\nu}$ of matter sources and their interaction with gravity (derived from $L^m$ according to the same rule (\ref{4})) participates on equal footing with $t^{\mu\nu}$ as the right hand side of the field equations: \begin{eqnarray} [(\gamma^{\mu\nu}+ h^{\mu\nu})(\gamma^{\alpha\beta}+ h^{\alpha\beta})- (\gamma^{\mu\alpha}+h^{\mu\alpha}) (\gamma^{\nu\beta}+h^{\nu\beta})]_{;\alpha;\beta} = \nonumber \\ =\frac{16 \pi G}{c^4}[t^{\mu\nu} + \tau^{\mu\nu}]. \label{14} \end{eqnarray} These equations are totally equivalent to the geometrical Einstein's equations \begin{equation} R^{\mu\nu}-\frac 1{2} g^{\mu\nu}R = \kappa T^{\mu\nu}, \end{equation} where $g^{\mu\nu}$ is defined by (\ref{g}), $R^{\mu\nu}$ is the Ricci tensor constructed from $g^{\mu\nu}$, and $T^{\mu\nu}$ is derived from the same matter Lagrangian (\ref{57}) according to the rule (\ref{1}). \section{Conclusion} The derived energy-momentum tensors $t^{\mu\nu}$ and $\tau^{\mu\nu}$ are fully satisfactory from the physical point of view and should be useful tools in practical applications. \section*{Acknowledgments} S. Babak was partially supported by ORS Award grant $N$ 97047008. \section*{References}
\section{Introduction} \label{secI} Structural properties of condensed matter depend sensitively on the space dimension $d$. Thin films offer the opportunity to reveal this dependence. By varying the film thickness $L$ one can interpolate smoothly between $d=2$ and $d=3$. For crystalline materials this variation can be accomplished with atomic resolution by using molecular beam epitaxy \cite{Herman:96}. As an alternative, which is applicable also for fluids, thin films can be built up via wetting phenomena where the film thickness is controlled by temperature or chemical potentials \cite{Dietrich:88}. Once such films are prepared the dependence of their structural properties on the space dimension can be studied particularly clearly close to phase transitions. For first-order phase transitions the main influence of a variation of the film thickness is to shift the phase boundaries in the phase diagram (see, e.g., capillary condensation \cite{Evans:90} or the shift of the melting curve \cite{Duffy:95}) without changing much the {\em local} properties of condensed matter. In rare cases, however, even the {\em character} of the phase transition can change as function of $L$; see, e.g., the possibility of continuous melting in $d=2$ \cite{Marcus:96} as opposed to $d=3$ or the crossover from a first-order phase transition in $d=3$ to a second-order phase transition in $d=2$ at a certain thickness of a slab of the 3-states Potts model \cite{An:88}. In the case of {\em first-order} phase transitions the robustness of the local structural properties with respect to changes of the film thickness is due to the smallness of the correlation lengths which characterize these systems and $-$ putting aside possible wetting phenomena $-$ thus severely limit the propagation of the structural changes, which necessarily occur near the confining surfaces of the film, into the interior of the films. In contrast, {\em second-order} phase transitions are characterized by diverging correlation lengths which affect not only the location of phase boundaries but in addition lead to pronounced changes in the local properties even deep in the interior of the films if the critical point is approached. These effects are thus not only particularly suitable to shed light on the aforementioned dependence of the structural properties on space dimension but they offer an additional advantage: the divergence of the correlation length as function of temperature upon approaching the critical point leads to {\em universal} behavior which makes a {\em quantitative} comparison between theoretical predictions and experiments much easier as compared with systems exhibiting first-order phase transitions which are characterized be several competing length scales of comparable, atomic size which are difficult to determine accurately and to vary systematically and independently. A sizable body of theoretical research has emerged describing continuous phase transitions in thin films (see, e.g., Refs. \cite{Fisher:70,Kaganov:72:a,Nauenberg:75,Suzuki:77,Berker:79,Dunfield:80,Brezin:82,Nightingale:82,Barber:83,Brezin:85,Rudnick:85}). Initiated by the theory of finite-size scaling (see, e.g., Refs. \cite{Binder:72,Fisher:72,Barber:73,Ritchie:73,Capehart:76,Callaway:81}), inter alia the shift $T_c(L)$ of the critical temperature with respect to its bulk value $T_c \equiv T_c(L = \infty)$ \cite{Allan:70,Bray:78,Nakanishi:83}, the magnetization \cite{Kaganov:72:b,Rudnick:82:a} as well as the free energy, the Casimir force, and the specific heat \cite{Krech:91,Krech:92:a,Eisenriegler:93,Krech:94,Sutter:94,Chen:95,Krech:96} have been analyzed. Here we emphasize that in order to observe universal film behavior the thicknesses $L$ of the films have still to be large on an atomic scale. This is assumed to be the case throughout our analysis. The analytic description of the dimensional crossover between $d=3$ critical behavior near $T_c$ and the $d=2$ critical behavior near $T_c(L)$ poses still a challenge \cite{Schmeltzer:85,OConnor:91:a} which has not yet been overcome with satisfactory quantitative accuracy. Numerous experiments (see, e.g., Refs. \cite{Gasparini:92,Ledermann:93,Nissen:93,Mehta:97,Andrieu:98,Henkel:98}) and simulations (see, e.g. Refs. \cite{Binder:88,Binder:95,Binder:97}) have been carried out to test these theoretical predictions. They lend support to the finite size scaling theory but still pose a puzzle as far as detailed quantitative agreement is concerned. The vast majority of these studies is devoted to integral or excess quantities without spatial resolution. However, the studies of {\em local} critical properties, such as of one- and two-point correlation functions, near a {\em single} surface have revealed a wealth of universal phenomena featuring numerous surface critical exponents and interesting crossover phenomena $-$ on the scale of the bulk correlation length $\xi$ $-$ between surface and bulk critical behavior \cite{Binder:83,Diehl:86}; the integral and excess quantities offer either no or only very limited access to these local properties. The successful development of surface specific X-ray and neutron scattering techniques based on exploiting total external reflection at grazing incidence has proven to be very fruitful, inter alia, for facilitating the quantitative comparison between experiments and theoretical predictions of the local critical behavior near interfaces \cite{Dietrich:95,Dosch:92}. These scattering techniques allow one to determine order parameter profiles normal to the surface and the depth resolved lateral two-point correlation function. In the present context such experiments have been carried out successfully for the binary alloy $Fe_3 Al$ \cite{Mailaender:90,Dosch:92:a,Dosch:92:b} and, by using truncation rod scattering, for $Fe Co$ \cite{Krimmel:97:b} which exhibit continuous order-disorder transitions in the bulk. In the case of $Fe_3 Al$ the cusplike surface singularities of the momentum and temperature dependence of the two-point correlation function turned out to be in excellent agreement with the theoretical predictions \cite{Dietrich:83:b,Dietrich:84}. The fact, that due to the occurrence of surface segregation suitable choices for the crystallographic orientation of the surface allows one to switch between the different surface universality classes corresponding to free boundary conditions and boundary conditions with surface fields, respectively, of the same bulk sample \cite{Drewitz:97,Leidl:98}, offers wide ranges of interesting comparative studies. In view of these developments and in view of the increasing availability of powerful synchrotron and neutron sources it appears promising to extend these studies of local critical properties to thin films. There are several predictions concerning the behavior of one-point correlation functions in thin films such as order-parameter profiles \cite{Kaganov:72:b,Rudnick:82:a,Krech:96} and energy-density profiles \cite{Eisenriegler:96}. However, on the level of the two-point correlation function so far only very little is known. This function depends on the lateral distance ${\bf x}_\| = {\bf x}^{(2)}_\| - {\bf x}^{(1)}_\|$ between the two points ${\bf x}_1 = ({\bf x}^{(1)}_\|, z_1)$ and ${\bf x}_2 = ({\bf x}^{(2)}_\|, z_2)$ (or equivalently the lateral momentum ${\bf p}$ corresponding to the $d-1$ translationally invariant directions), the coordinates $z_1$ and $z_2$ perpendicular to the parallel surfaces of the film, the film thickness $L$, and temperature $t = (T-T_c)/T_c$ (or equivalently the bulk correlation length $\xi = \xi_0 t^{-\nu}$). Since a full sweep of this large parameter space is practically not possible for computer simulations, we have applied fieldtheoretic techniques which provide analytic access to the full parameter space. This approach encompasses nonperturbative features such as scaling properties and short-distance expansions as well as an explicit and systematic perturbative result to first order in $\epsilon = 4 - d$. The latter serves as to corroborate the nonperturbative results and to provide numerical results which are not accessible by general arguments. These explicit calculations are carried out for the fixed point of the so-called ordinary transition for both confining surfaces in the classification scheme of surface critical phenomena \cite{Diehl:86} corresponding to free boundary conditions on both sides. This is applicable to thin antiferromagnetic films near their N\'eel temperature, to ferromagnetic films near their Curie temperature in the absence of external bulk and surface fields, and to thin films of binary alloys near their continuous order-disorder transitions. Among the numerous order-disorder phase transitions in binary alloys only a few are of second order including $Fe_3 Al$ \cite{Guttmann:56,Guttmann:69:a,Guttmann:69:b}, $FeCo$ \cite{Oyedele:77}, $CuZn$ \cite{AlsNielsen:67}, and $FeAl$ \cite{Guttmann:69:a}. Both the $B2-DO_3$ transition in $Fe_3Al$ and the $A2-B2$ transitions in $FeCo$, $CuZn$, and $FeAl$ belong to the Ising universality class \cite{Diehl:97}. For the $A2-B2$ transitions it is predicted theoretically that the $(110)$ surface belongs to the surface universality class of the ordinary transition whereas the $(100)$ surface exhibits the so-called normal transition associated with the presence of an effective surface field \cite{Drewitz:97,Leidl:98}. Indeed truncation rod scattering at the $FeCo$ $(100)$ surface has provided clear evidence for the presence of an effective surface field \cite{Krimmel:97:b} above $T_c$ although the expected associated crossover from ordinary to normal critical behavior \cite{Ritschel:96} could not yet been resolved experimentally in an unequivocal way. The results of the diffuse scattering of X-rays under grazing incidence from the $(1 \bar 1 0)$ surface (equivalent to the $(110)$ surface) of $Fe_3Al$ \cite{Mailaender:90} are in excellent agreement with the theoretical predictions \cite{Dietrich:83:b,Dietrich:84} for the ordinary transition. But even for $Fe_3Al$ $(1 \bar 1 0)$ a residual order parameter above $T_c$ has been reported \cite{Mailaender:90,Dosch:92:a}. Thus it still remains to be seen theoretically whether for the $B2-DO_3$ transition in $Fe_3Al$, in contrast to the $A2-B2$ transition in $FeCo$, the $(1 \bar 1 0)$ surface can support a weak effective surface field. In view of this state of affairs our present result are expected to be closely applicable to thin films of $Fe_3Al$, $FeCo$, $CuZn$, and $FeAl$ bounded by $(110)$ surfaces on both sides. Among them $Fe_3Al$ and $FeCo$ appear to be the most promising candidates because the others exhibit strong surface segregation. For an assessment of the possibilities to probe critical magnetic surface transitions by grazing incidence of neutrons see Ref. \cite{Dosch:93}. In view of the aforementioned difficulties concerning the analytic description of the dimensional crossover we confine our analysis to the temperature range $T \ge T_c$. We note that elements of the perturbation theory for thin critical films can be found in Ref. \cite{Nemirovsky:86:a}. We had, however, to carry out our own approach because the representation given in Ref. \cite{Nemirovsky:86:a} is not suited for making predictions for the scattering experiments and because Ref. \cite{Nemirovsky:86:a} contains errors. Finally we note that experience tells that calculations carried out for the spherical model, as have been done for the present system \cite{Allen:94}, lack the quantitative reliability needed for comparison with experiments and simulations. In order to encourage future scattering experiments for critical thin films and to facilitate an explicit quantitative comparison of such data with the present theoretical predictions, we have calculated the singular contributions to the scattering cross section for X-ray and neutron scattering under the condition of grazing incidence based on our results for the critical two-point correlation function in thin films. This allows us to describe the conditions under which the various singularities of the two-point correlation function become visible in scattering data. This introduction is followed be three sections, the Summary, and four Appendices. In Sec. \ref{model} we introduce the fieldtheoretical model. The two-point correlation function is discussed in Sec. \ref{correlation} and in Sec. \ref{crosssection} we investigate the scattering cross section. Relations between bulk and film amplitudes are derived in Appendix \ref{amplitudes}, explicit one-loop results are presented in Appendix \ref{one_loop_results}, and Appendices \ref{details_crosssection} and \ref{ssss} contain details required for the calculation of the scattering cross section. \section{Field-theoretical model} \label{model} The leading critical behavior in a film follows from the statistical weight $\exp (-{\cal H} \{ \Phi \})$ for the configuration $\Phi ({\bf x}) = (\phi_a ({\bf x}), a=1, \dots , n)$ of a $n$-component field, which is proportional to the order parameter, where \cite{Diehl:86,Krech:92:a,Krech:94} \begin{eqnarray} \label{free_energy} {\cal H} \{ {\Phi} ({\bf x}) \} & = & \int d^{d-1}x_\Vert \int_0^L dz \left( \frac{1}{2} (\nabla \Phi)^2 + \frac{\tau}{2} \Phi^2 + \frac{g}{4!} (\Phi^2)^2 - {\bf h} \cdot \Phi \right) \\ &+& \int d^{d-1}x_\Vert \left( \frac{c}{2} \Phi^2 (z=0) - {\bf h}_1 \cdot \Phi(z=0) +\frac{c}{2} \Phi^2 (z=L) - {\bf h}_1 \cdot \Phi(z=L) \right) \nonumber \end{eqnarray} with space dimension $d$ and position vector ${\bf x}=({\bf x_\Vert},z)$ of $d-1$ parallel and one perpendicular components. The $z$ integration extends over the interval $[0,L]$, where $z=0$ and $z=L$ give the positions of the film surfaces. $\tau$ is the temperature parameter such that in the bulk $\tau = 0$ marks the transition temperature within mean-field theory. The coupling constant $g>0$ ensures the stability of the statistical weight below the transition temperature, i.e., for $\tau < 0$. $c$ denotes the surface enhancement, ${\bf h}$ and ${\bf h}_1$ are bulk and surface fields, respectively. We focus on the ordinary transition at zero fields, i.e., we adopt the fixed point value $c=\infty$ for the surface enhancement and set ${\bf h}={\bf h}_1=0$. After carrying out a Fourier transformation with respect to the $d-1$ directions exhibiting translational invariance parallel to the surfaces the mean-field propagator for the disordered phase $(\tau > 0)$ in $p$-$z$-representation is given by \cite{Diehl:86,Krech:95} \begin{eqnarray} \label{propagator_dirichlet} G_D(p,z_1,z_2,L,\tau) & = & \int d^{d-1}x_\| \enspace e^{i {\bf p} \cdot {\bf x}_\|} \enspace \langle \Phi ({\bf x}_\|,z_1) \Phi (0,z_2 ) \rangle \\ & = & \frac{1}{2 b} \Big( e^{-b|z_1-z_2|} - e^{-b(z_1+z_2)} \nonumber \\ && \quad + \left. \frac{e^{-b(z_1-z_2)} + e^{-b(z_2-z_1)} - e^{-b(z_1+z_2)} - e^{b(z_1+z_2)}}{e^{2bL}-1} \right) , \enspace b=\sqrt{p^2+\tau}. \nonumber \end{eqnarray} The first exponential function corresponds to the bulk part followed by the contribution from the surface at $z=0$. Both exponentials together give the propagator for the ordinary transition of the semi-infinite system $(L=\infty)$. The remaining ratio carries the $L$ dependence. The propagator satisfies the Dirichlet boundary conditions $G_D(z=0)=0=G_D(z=L)$. Equation (\ref{propagator_dirichlet}) represents the mean-field approximation for the two-point correlation function in the film corresponding to the critical behavior in $d=4$. The non-Gaussian fluctuations in $d=3$ are taken into account approximately by the one-loop correction which amounts to the first term in a systematic expansion in terms of $\epsilon = 4 - d$: \begin{eqnarray} \label{one_loop} &&G_{bare}(p,z_1,z_2,L,\tau,g) = G_D(p,z_1,z_2,L,\tau) \\ & - & \frac{g}{2} \frac{n+2}{3} \int \frac{d^{d-1}q}{(2\pi)^{d-1}} \int\limits_0^L dz \enspace G_D(p,z_1,z,L,\tau) \enspace G_D(q,z,z,L,\tau) \enspace G_D(p,z,z_2,L,\tau) + {\cal O}(g^2). \nonumber \end{eqnarray} As regularization scheme we use dimensional regularization by analytic continuation in the space dimension $d=4-\epsilon $. As long as $z_1$ and $z_2$ are both off the surfaces only bulk singularities occur. We absorb the corresponding poles in $\epsilon$ by minimal subtraction through the standard $Z$ factors: \begin{eqnarray} \label{renormalized_parameters} \phi = Z^{1/2}_{\phi} \phi^R , \quad g = \mu^{\epsilon} 2^d \pi^{d/2} Z_u u , \quad \tau = \mu^2 Z_t t , \end{eqnarray} where $\mu$ is the momentum scale and the bulk Z-factors are \cite{Amit:78} \begin{eqnarray} \label{bulk_z_factors} Z_{\phi} = 1 + {\cal O} (u^2), \quad Z_u = 1 + \frac{n+8}{3} \frac{u}{\epsilon} + {\cal O} (u^2), \quad Z_t = 1 + \frac{n+2}{3} \frac{u}{\epsilon} + {\cal O} (u^2). \end{eqnarray} The renormalized correlation function reads (see Eq. (\ref{renormalized_parameters})) \begin{eqnarray} \label{renormalization} G(p,z_1,z_2,L,t,u;\mu) & = & Z_{\phi}^{-1} G_{bare}(p,z_1,z_2,L,\tau,g) \end{eqnarray} which is valid in all orders of perturbation theory. The solution of the corresponding renormalization group equation leads to the following scaling property: \begin{eqnarray} \label{scaling_form} G(p,z_1,z_2,L,t;\mu) & = & {\cal G}_{\mbox{\tiny I}} p^{-1+\eta} g_{\mbox{\tiny I}} (p \xi, z_1/\xi, z_2/\xi, L/\xi) . \end{eqnarray} This holds at the fixed point $u^\ast = \frac{3}{n+8}\epsilon + {\cal O} (\epsilon^2)$ and involves the bulk correlation length $\xi=\xi^+_0 t^{-\nu}$, the exponents $\eta = {\cal O} (\epsilon^2)$, and $\nu = \frac{1}{2} + \frac{1}{4} \frac{n+2}{n+8} \epsilon + {\cal O} (\epsilon^2)$. With suitable normalization (see, c.f., Eq. (\ref{norm_ri})) the scaling function $g_{\mbox{\tiny I}}$ is universal. The amplitude ${\cal G}_{\mbox{\tiny I}}$, which is fixed by this normalization, and the amplitude $\xi^+_0$ carry the nonuniversal scaling factors. We fix $\xi^+_0$ by defining $\xi$ as the so-called true correlation length \cite{Tarko:75} so that $\xi^+_0 = \mu^{-1} (1+ \frac{1}{4} \frac{n+2}{n+8} (1-C_E) \epsilon + {\cal O} (\epsilon^2))$. This expression for $\xi^+_0$ allows one to express the momentum scale $\mu$ introduced in Eq. (\ref{renormalized_parameters}) in terms of the experimentally accessible, nonuniversal amplitude $\xi^+_0$: \begin{eqnarray} \label{mu_fixing} \mu & = & (\xi^+_0)^{-1} \Big( 1 + \frac{1}{4} \frac{n+2}{n+8} (1-C_E) \epsilon + {\cal O} (\epsilon^2) \Big) . \end{eqnarray} If subsequent formulae contain the momentum scale $\mu$ explicitly it is to be replaced by Eq. (\ref{mu_fixing}); moreover we omit $\mu$ from the explicit list of variables of $G$. Depending on the problem under consideration it is often advantageous to use different but equivalent representations of the correlation function such as \begin{eqnarray} \label{alternatives_1} G(p,z_1,z_2,L,t) & = & {\cal G}_{\mbox{\tiny II}} z_1^{1-\eta} g_{\mbox{\tiny II}} (p z_2, z_1/\xi, z_2/\xi, z_2/L) , \end{eqnarray} \begin{eqnarray} \label{alternatives_2} G(p,z_1,z_2,L,t) & = & {\cal G}_{\mbox{\tiny III}} L^{1-\eta} g_{\mbox{\tiny III}} (p z_1, z_1/L, z_2/L, L/\xi) , \end{eqnarray} \begin{eqnarray} \label{alternatives_3} G(p,z_1,z_2,L,t) & = & {\cal G}_{\mbox{\tiny IV}} \xi^{1-\eta} g_{\mbox{\tiny IV}} (p L, p z_1, p z_2, \xi/L) , \end{eqnarray} and \begin{eqnarray} \label{alternatives_4} G(p,z_1,z_2,L,t) & = & {\cal G}_{\mbox{\tiny V}} p^{-1+\eta} g_{\mbox{\tiny V}} (p \xi, p (z_1-z_2), p (z_1+z_2), L/\xi) . \end{eqnarray} The nonuniversal amplitudes ${\cal G}_x$ and the universal scaling functions $g_x$, $x={\mbox{\tt I,II,III,IV,V}}$, are fixed by the following normalizations: \begin{eqnarray} \label{norm_ri} \lim_{\alpha \to \infty} \lim_{\beta \to \infty} \lim_{\delta \to \infty} g_{\mbox{\tiny I}} (\alpha,\beta,\gamma=\beta,\delta) =1, \end{eqnarray} \begin{eqnarray} \label{norm_rii} \lim_{\alpha \to 0} \lim_{\beta \to 0} \lim_{\delta \to 0} g_{\mbox{\tiny II}}(\alpha,\beta,\gamma = \beta,\delta) =: g_{\mbox{\tiny II}}(0,0,0,0)=1, \end{eqnarray} \begin{eqnarray} \label{norm_riii} \lim_{\alpha \to 0} \lim_{\delta \to 0} g_{\mbox{\tiny III}} (\alpha,\beta=1/2,\gamma=\beta=1/2,\delta) =1, \end{eqnarray} \begin{eqnarray} \label{norm_riv} \lim_{\delta \to 0} \lim_{\alpha \to 0} g_{\mbox{\tiny IV}}(\alpha,\beta=\alpha/2,\gamma=\beta=\alpha/2,\delta) =1, \end{eqnarray} and \begin{eqnarray} \label{norm_rv} \lim_{\beta \to 0} \lim_{\alpha \to \infty} \lim_{\gamma \to \infty} \lim_{\delta \to \infty} g_{\mbox{\tiny V}}(\alpha,\beta,\gamma,\delta) =: g_{\mbox{\tiny V}}(\infty,0,\infty,\infty)=1. \end{eqnarray} The universal scaling functions $g_x$ can be expressed in terms of each other because in Eqs. (\ref{scaling_form}) and (\ref{alternatives_1}) - (\ref{alternatives_4}) the left hand side is the same quantity and the sets of scaling variables are complete, i.e., from each set one can form any of the others by a suitable combination of variables. Since the nonuniversal amplitudes ${\cal G}_x$ correspond to the same correlation function $G(p,z_1,z_2,L,t)$ and because the scaling functions fixed by the normalizations in Eqs. (\ref{norm_ri}) - (\ref{norm_rv}) are universal, their ratios ${\cal G}_x/{\cal G}_{x'}$ are universal numbers. Thus the knowledge of one of them and of the corresponding universal scaling functions determines all the others. Moreover, as discussed in Appendix \ref{amplitudes}, all nonuniversal amplitudes ${\cal G}_x$ are determined by any pair of nonuniversal scale factors which characterize the critical {\em bulk} properties. A transparent and experimentally directly accessible choice for the latter is the nonuniversal amplitude $B$ of the leading temperature singularity of the field $\langle \phi ({\bf x}) \rangle$ in the bulk below $T_c$, \begin{eqnarray} \label{bulk_op} \langle \phi ({\bf x}) \rangle & = & B (-t)^\beta , \end{eqnarray} and the amplitude $\xi_0^+$ of the true correlation length above $T_c$. In terms of these quantities one has \begin{eqnarray} \label{G_e_B_xi0} {\cal G}_{\mbox{\tiny V}} & = & B^2 (\xi_0^+)^{d-2+\eta} {\cal U} \end{eqnarray} where ${\cal U}$ is a universal number, whose value ${\cal U} \simeq 1.58$ is derived in Appendix \ref{amplitudes} based on Eq. (\ref{norm_rv}). In the following most of our analysis focuses on the scaling function $g_{\mbox{\tiny II}}$ used in Eq. (\ref{alternatives_1}). For that case one finds (see Appendix \ref{amplitudes}) the {\em universal} ratio \begin{eqnarray} \label{G_e_G_b} {\cal G}_{\mbox{\tiny II}}/{\cal G}_{\mbox{\tiny V}} & = & 2 \Big( 1 + \epsilon \frac{n+2}{n+8} + {\cal O} (\epsilon^2) \Big) . \end{eqnarray} With these results we finally obtain \begin{eqnarray} \label{G_final} G(p,z_1,z_2,L,t) & = & B^2 (\xi_0^+)^{d-1} {\cal R} (z_1/\xi_0^+)^{1-\eta} g_{\mbox{\tiny II}} (p z_2, z_1/\xi, z_2/\xi, z_2/L) \end{eqnarray} where ${\cal R} = 2 {\cal U} \Big( 1 + \epsilon \frac{n+2}{n+8} + {\cal O} (\epsilon^2) \Big) \simeq 4.21$ is a universal number. Thus in all our subsequent formulae for {\em film} properties their {\em absolute} values are determined and fixed by the two nonuniversal {\em bulk} amplitudes $B$ and $\xi_0^+$. The actual order parameter $O\!P$ for a particular second order phase transition is proportional to the field $\phi$ introduced in Eq. (\ref{free_energy}), i.e., $O\!P({\bf x}) = {\it b} \phi ({\bf x})$. The value of ${\it b}$ depends on the particular system (binary alloy, liquid, ferromagnet etc.). Moreover, any rescaling of ${\it b}$ by a dimensionless number renders another order parameter $O\!P$ which is equally valid for describing the singular behavior of the phase transition. We emphasize that Eqs. (\ref{alternatives_1}), (\ref{G_e_B_xi0}), and (\ref{G_e_G_b}) remain valid if $G$ is replaced by $\langle O\!P ({\bf x}) O\!P ({\bf x}') \rangle$, $\langle \phi ({\bf x}) \rangle$ by $\langle O\!P ({\bf x}) \rangle$, and $B$ by $B'= {\it b} B$; these replacements have to be carried out if the present fieldtheoretic results are used to interpret, e.g., the intensity of scattered X-rays or neutrons (see, c.f., Sec. \ref{crosssection}). The actual choice of the $O\!P$, as it enters into the expression for the scattering cross section, is borne out and tight to the relation $\langle O\!P ({\bf x}) \rangle = B' (-t)^\beta $. \section{Explicit properties of the two-point correlation function} \label{correlation} The discussion of the correlation function consists of three parts. First, we set $z_1=z_2$ and analyze its nonanalytic behavior in certain limits. Then, we take into account the case $z_1 \not= z_2$, which serves to understand the correlations perpendicular to the surfaces. Moreover, the discussion of this latter case turns out to be very useful for carrying out the integrations appearing in the scattering cross section to be analyzed in Sec. \ref{crosssection}. The film excess susceptibility is discussed in the last part. \subsection{Lateral two-point correlation function for $z_1=z_2$} \label{correlation_zz} In order to investigate various asymptotic properties of the lateral behavior of the two-point correlation function we resort to short distance expansions (SDE) \cite{Diehl:81:a}, distant wall corrections (DWC) \cite{Eisenriegler:96}, and results of the perturbation theory supported by appropriate exponentiations of the explicit $\epsilon$-expansion results. With $z_1=z_2=z$, in the present context a representation of the form \begin{eqnarray} \label{corr_fctn_z1_eq_z2} G(p,z,L,t) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} g(pz,z/\xi,z/L) \end{eqnarray} is useful. According to Eq. (\ref{alternatives_1}) one has $g(u,v,w)=g_{\mbox{\tiny II}}(u,v,v,w)$ with $g(0,0,0)=1$ (Eq. (\ref{norm_rii})). For semi-infinite systems, i.e., $L=\infty$ the SDE in the cases $t=0$, $p \to 0$ and $p=0$, $t \to 0$ \cite{Gompper:84,Gompper:86:a} leads to the asymptotic behaviors \begin{eqnarray} \label{corr_p} G(p,z,L=\infty,t=0) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} g_1(u=pz) \\ & _{\longrightarrow \atop p \to 0} & {\cal G}_{\mbox{\tiny II}} z^{1-\eta}[1 + A_1 (pz)^{-1+\eta_\|} + \dots] \nonumber \end{eqnarray} and \begin{eqnarray} \label{corr_t} G(p=0,z,L=\infty,t) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta}g_2(v=z/\xi) \\ & _{\longrightarrow \atop t \to 0} & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} [1 + B_1 (z/\xi)^{-1+\eta_\|} + \dots] \nonumber \\ & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} [1 + B_1 (z/\xi^+_0)^{-1+\eta_\|} t^{-\gamma_{11}} + \dots], \nonumber \end{eqnarray} respectively, with $\gamma_{11}=\nu (\eta_\|-1)$, $g_1(u) = g(u,v=0,w=0)$, $g_1(0)=1$, $g_2(v) = g(u=0,v,w=0)$, and $g_2(0)=1$. In the case $p=0$, $t=0$ one has \begin{eqnarray} \label{corr_L_1} G(p=0,z,L,t=0) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} g_3(w=z/L) \end{eqnarray} with $g_3(w)=g(u=0,v=0,w)$ so that $g_3(0)=1$. In order to infer the first nontrivial dependence on $L$ for $L \to \infty$, according to Eq. (\ref{corr_L_1}) one can equally consider the limit $z \to 0$ for $L$ fixed. To this end we consider the SDE of the renormalized film correlation function in real space: \begin{eqnarray} \label{sde_dwc} \langle \phi({\bf x}_\|,z) \phi(0,z) \rangle & _{\longrightarrow \atop z \to 0} & \mu^{-2} (\mu z)^{2(x_s-x)} \langle \phi_\perp({\bf x}_\|) \phi_\perp(0) \rangle \\ & = & \mu^{d-2} (\mu z)^{2(x_s-x)} (\mu x_\|)^{-2x_s} Y(x_\|/L) \nonumber. \end{eqnarray} Here $\phi_\perp$ denotes the normal derivate of $\phi$ taken at one of the surfaces and $Y(y)$ is a dimensionless scaling function for the film which is universal up to nonuniversal prefactor. The scaling dimensions of $\phi$ and $\phi_\perp$ are $x=\frac{1}{2}(d-2+\eta)$ and $x_s=\frac{1}{2}(d-2+\eta_\|)$, respectively. The scaling function $Y(x_\|/L)$ describes the influence of the distant wall at $z=L$ on the lateral correlations close to the near wall at $z=0$. In order to obtain its leading asymptotic behavior for $x_\|/L \to 0$ we use the identity \begin{eqnarray} \label{G_L_identity} G(p=0,z,L,t=0) & = & G(p=0,z,L=\infty,t=0) \\ && - \int_L^\infty \frac{\partial G(p=0,z,L',t=0)}{\partial L'} d L' . \nonumber \end{eqnarray} The first term on the rhs is equal to ${\cal G}_{\mbox{\tiny II}} z^{1-\eta}$ (compare Eqs. (\ref{corr_p}) and (\ref{corr_t})). The leading correction is given by using the SDE in Eq. (\ref{sde_dwc}) for the second term: \begin{eqnarray} \label{leading_L_correction} -\int\limits_L^\infty \frac{\partial G(p=0,z,L',t=0)}{\partial L'} d L'&=&- \int d^{d-1} x_\| \int\limits_L^\infty d L' \frac{\partial}{\partial L'} \langle \phi({\bf x}_\|,z) \phi(0,z) \rangle \\ & _{\longrightarrow \atop L \to \infty} & - \int d^{d-1} x_\| \int\limits_L^\infty d L' \frac{\partial}{\partial L'} \mu^{d-2} (\mu z)^{2(x_s-x)} (\mu x_\|)^{-2x_s} Y(x_\|/L') \nonumber \\ & = & \mu^{-1} (\mu z)^{1-\eta} \Big( \frac{z}{L} \Big)^{-1+\eta_\|}\tilde C \nonumber \end{eqnarray} with $\tilde C = \frac{1}{\eta_\|-1} \int d^{d-1}y \enspace y^{-(d-3+\eta_\|)} Y'(y)$. Thus we find $g_3(w \to 0) = 1 + C_1 w^{-1+\eta_\|}$ where $C_1 = \tilde C \mu^{-\eta}/{\cal G}_{\mbox{\tiny II}}$ is a universal number, i.e., \begin{eqnarray} \label{corr_L} G(p=0,z,L \to \infty,t=0) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta}[1 + C_1 (z/L)^{-1+\eta_\|} + \dots] . \end{eqnarray} Finally we note that due to the normalization $g(0,0,0)=1$ the scaling function $g(u,v,w)$ is given by the ratio $g(pz,z/\xi,z/L)=G(p,z,L,t)/G(p=0,z,L=\infty,t=0)$ from which the prefactors ${\cal G}_{\mbox{\tiny II}} z^{1-\eta}$ appearing in Eq. (\ref{corr_fctn_z1_eq_z2}) drop out. The $\epsilon$-expansions of the amplitudes of the leading asymptotic terms follow from Eqs. (\ref{g_p_to_0}), (\ref{g_t_to_0}), and (\ref{g_L_to_oo})) in Appendix \ref{zz_appendix}: \begin{eqnarray} \label{A1_B1_C1} A_1 & = & - \Big[ 1 + \epsilon \frac{n+2}{n+8} \left( 1 - C_E - \ln 2 \right) + {\cal O} (\epsilon^2) \Big] , \\ B_1 & = & - \Big[ 1 + \epsilon \frac{n+2}{n+8} \left( 1 - C_E \right) + {\cal O} (\epsilon^2) \Big] , \nonumber \\ C_1 & = & - \Big[ 1 + \epsilon \frac{n+2}{n+8} \left( \frac{\pi^2}{18} - C_E +2(S_2+I_1) - 1\right) + {\cal O} (\epsilon^2) \Big] . \nonumber \end{eqnarray} $C_E \approx 0.5772$ is Euler's constant, $S_2 \simeq 0.083$ and $I_1 \simeq 0.287$ are given by Eq. (\ref{S2_I1}) in Appendix \ref{zz_appendix}. Within the $\epsilon$-expansion the full forms of the scaling functions $g_1(u)$, $g_2(v)$, and $g_3(w)$ can be found in Appendix \ref{zz_appendix} (see Eqs. (\ref{g_p}) - (\ref{g_L})). In Fig. \ref{fig_p_t_L} we display the three scaling functions $g_i$, $i=1,2,3$, (Eqs. (\ref{g_p}) - (\ref{g_L})) corresponding to Eqs. (\ref{corr_p}), (\ref{corr_t}), and (\ref{corr_L_1}) as obtained within mean-field theory (MFT), i.e., for $\epsilon=0$ and from renormalization group guided perturbation theory (PT) as well as their leading behavior $g_i (x_i \to 0) = g_{i,l}(x_i)$, $x_1=u, x_2=v, x_3 =w$. Within MFT the three scaling functions have the same limiting form for small scaling variables with $A_1=B_1=C_1=-1$ and the critical exponent $\eta_\|=2$. Beyond MFT, in Fig. \ref{fig_p_t_L} we use $\eta_\|=1.48$ as the best available estimate \cite{Diehl:86} whereas the amplitudes are evaluated in first order in $\epsilon$ (Eq. (\ref{A1_B1_C1}) for $(n,\epsilon)=(1,1)$) so that $A_1 \simeq -0.9099$, $B_1 \simeq -1.1409$, and $C_1 \simeq -0.9035$. Within mean-field theory $g_1=g_2$ and the leading asymptotic behavior $g_{3,l}$ provides already the full scaling function $g_3$. Beyond MFT there is a small difference between $g_3$ and $g_{3,l}$. This difference is much bigger for the scaling functions $g_1$ and $g_2$ describing the semi-infinite system. The above discussion demonstrates that, for $z$ fixed, the two-point correlation function $G(p,z,L,t)$ has a finite value $G(p=0,z,L=\infty,t=0)$ which is attained via cusplike singularities: $\sim p^{-1+\eta_\|} \enspace (p \to 0, 1/L =0, t=0)$, $\sim (1/L)^{-1+\eta_\|} \enspace (1/L \to 0, p=0, t=0)$, $\sim (1/\xi)^{-1+\eta_\|} \enspace (t \to 0, p=0, 1/L =0)$. In terms of these variables the critical exponent is the same for all three cases and only the amplitudes differ. These singularities remain if only one out of the above three variables is zero and the remaining two both vanish. This behavior, which includes the smooth interpolation between the corresponding amplitudes, is described by the scaling functions $h_1$, $h_2$, and $h_3$ of two variables instead of the scaling functions with one variable as $g_1(u)$, $g_2(v)$, and $g_3(w)$: \begin{eqnarray} \label{g_pt} G(p,z,L=\infty,t) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} h_1(u,v), \enspace h_1(u,v) = g(u,v,w=0) , \end{eqnarray} \begin{eqnarray} \label{g_pL} G(p,z,L,t=0) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} h_2(u,w), \enspace h_2(u,w) = g(u,v=0,w) , \end{eqnarray} and \begin{eqnarray} \label{g_tL} G(p=0,z,L,t) & = & {\cal G}_{\mbox{\tiny II}} z^{1-\eta} h_3(v,w), \enspace h_3(v,w) = g(u=0,v,w) \end{eqnarray} with $u=pz$, $v=z/\xi$, and $w=z/L$. All three scaling function can be obtained from Eq. (\ref{g_ptL_z1z2}). Since the discussion of all three scaling functions is analogous we demonstrate our analysis only for $h_3(v,w)$. We introduce polar coordinates $\omega$ and $\varphi$ \begin{eqnarray} \label{polar_coordinates} \omega & = & \sqrt{v^2+w^2} = z \sqrt{\xi^{-2}+L^{-2}}, \enspace \varphi = \arctan(v/w) = \arctan (L/\xi) \enspace , \\ v & = & \omega \sin \varphi, \enspace w = \omega \cos \varphi \nonumber \end{eqnarray} which leads to \begin{eqnarray} \label{h3_transf} h_3(v,w) & = & h_3 (\omega \sin \varphi, \omega \cos \varphi) = h^{(3)}_{polar} (\omega,\varphi) . \end{eqnarray} Since the limit $\omega \to 0$, i.e., $1/\xi \to 0$ and $1/L \to 0$, is equivalent to the limit $z \to 0$ for $\xi$ and $L$ fixed the resulting singularity is compatible with the SDE so that \begin{eqnarray} \label{h3_polar_0} h^{(3)}_{polar} (\omega \to 0, \varphi) & = & H^{(3)}_0(\varphi) + H^{(3)}_1(\varphi) \omega^{-1+\eta_\|} + \dots \enspace . \end{eqnarray} The explicit form of the scaling function $h_3(v,w)$ as obtained from perturbation theory in ${\cal O}(\epsilon)$ is in accordance with Eq. (\ref{h3_polar_0}) and renders explicit results for the coefficients $H^{(3)}_0(\varphi)$ and $H^{(3)}_1(\varphi)$: \begin{eqnarray} \label{H_0_phi} H_0(\varphi) & = & h^{(3)}_{polar} (\omega=0,\varphi) = h_3(v=0,w=0) = 1 \end{eqnarray} is independent of $\varphi$ and equal to 1 due to the normalization $g(u=0,v=0,w=0)=1$. With this result the $\epsilon$-expansion of $H^{(3)}_1(\varphi)$ follows by comparing the $\epsilon$-expansion of the rhs of Eq. (\ref{h3_polar_0}) with the limit $\omega \to 0$ of the $\epsilon$-expansion of $h^{(3)}_{polar}(\omega,\varphi)$. As expected one finds that $H^{(3)}_1(\varphi)$ interpolates smoothly between the value $H^{(3)}_1(\varphi=0)=C_1$ (see Eq. (\ref{A1_B1_C1})) corresponding to the amplitude of the singularity $\sim (1/L)^{-1+\eta_\|}$ for $u=0$ and $v=0$ and the value $H^{(3)}_1(\varphi=\pi/2)=B_1$ (see Eq. (\ref{A1_B1_C1})) corresponding to the amplitude of the singularity $\sim (1/\xi)^{-1+\eta_\|}$ for $u=0$ and $w=0$. In Fig. \ref{fig_H1H2H3_phi} all three amplitude functions $H^{(1)}_1(\varphi)$, $H^{(2)}_1(\varphi)$, and $H^{(3)}_1(\varphi)$ (see Eqs. (\ref{H1_1}) - (\ref{H1_3})) are shown in mean-field theory (MFT) and in first order in $\epsilon$ (PT). Within MFT $H^{(1)}_1(\varphi)$ of the semi-infinite system is constant and $H^{(2)}_1(\varphi) = H^{(3)}_1(\varphi)$ exhibit a nontrivial dependence on $\varphi$. Beyond MFT all three functions interpolate between the amplitudes $A_1$, $B_1$, and $C_1$ (see Eq. (\ref{A1_B1_C1})) in a nontrivial way. In Figs. \ref{fig_cusp_pt}, \ref{fig_cusp_pL}, and \ref{fig_cusp_tL} we display the full scaling functions $h_1(u,v)$, $h_2(u,w)$, and $h_3(v,w)$, respectively. In order to obtain such a scaling function beyond the leading asymptotic form we first subtract its leading contribution in its $\epsilon$-expanded form in ${\cal O} (\epsilon)$ from the full expression of the scaling function and add the leading exponentiated contribution afterwards. This exponentiation scheme is consistent with the explicit expanded form up to and including ${\cal O} (\epsilon)$. In Figs. \ref{fig_p_cusp_tL}, \ref{fig_t_cusp_pL}, and \ref{fig_L_cusp_pt} we show cross sections of the three-dimensional plots in order to illustrate the emergence of the $p^{-1+\eta_\|}$ cusplike singularity upon varying $t$ or $L$, the $(1/\xi)^{-1+\eta_\|}$ cusplike singularity upon varying $p$ or $L$, and the $(1/L)^{-1+\eta_\|}$ cusplike singularity upon varying $t$ or $p$, respectively. \subsection{Perpendicular correlations} \label{correlation_z1z2} In a semi-infinite system the perpendicular correlations in real space define the exponent $\eta_\perp = (\eta + \eta_\|)/2$ through the limit $G(x_\|,z_1 \to \infty,z_2,L=\infty,t) \sim z_1^{-(d-2+\eta_\perp)}$ with $x_\|$ and $z_2$ fixed. A Fourier transformation leads to the relation $G(p=0,z_1,z_2,L=\infty,t) \sim z_1^{1-\eta_\perp}$ with $z_2$ fixed and $z_1 \to \infty$. Note that in real space $G(x_\|,z_1,z_2,L=\infty,t=0)$ {\em increases} as function of $z_1$ for $z_2$ and $x_\|$ fixed, reaches a maximum at a certain value $z_1^\ast =z_2 f(x_\|/z_2)$ and finally vanishes for $z_1 \to \infty$. This increase for $0 < z_1 < z_1^\ast$ leads to the divergence $\sim z_1^{1-\eta_\perp}$, $1-\eta_\perp \simeq 0.25$, of $G(p=0,z_1,z_2,L=\infty,t=0) = \int d x_\| \langle \phi (0,z_1) \phi (x_\|,z_2) \rangle$. The coordinates $z_1$ and $z_2$ can be interchanged. Actually conformal invariance fixes completely the functional form of $G(p=0,z_1,z_2,L=\infty,t=0)$ (see Ref. \cite{Cardy:84}). SDE leads up to a constant amplitude to the expression \begin{eqnarray} \label{gompper_cft_z1z2} G(p=0,z_1,z_2,L=\infty,t=0) & \sim & (z_1 z_2)^\frac{1-\eta}{2} \Big ( \Theta(z_1-z_2) \frac{z_2}{z_1} + \Theta(z_2-z_1) \frac{z_1}{z_2} \Big)^\frac{\eta_\|-1}{2} + \dots \end{eqnarray} (see Eq. (4.68) in Ref. \cite{Gompper:86:b}). The explicit calculation to first order in $\epsilon$ gives \begin{eqnarray} \label{gompper_firstorder_z1z2} G(p=0,z_1,z_2,L=\infty,t=0) & = & {\cal G}_{\mbox{\tiny II}} (z_1 z_2)^\frac{1-\eta}{2} \Big ( \Theta(z_1-z_2) \frac{z_2}{z_1} + \Theta(z_2-z_1) \frac{z_1}{z_2} \Big)^\frac{\eta_\|-1}{2} \end{eqnarray} for arbitrary $z_1$ and $z_2$. This perturbation theory guided result for $d=3$ has a structure similar to the exact result from conformal theory in $d=2$ (see Refs. \cite{Cardy:84} and \cite{Gompper:86:b}). Therefore one is led to the conclusion that Eq. (\ref{gompper_firstorder_z1z2}) is a good approximation for the exact correlation function in $d=3$. Guided by these considerations we find that in the case that the variables $p$, $t$, and $1/L$ are small but nonzero the explicit results for $G$ obtained from the $\epsilon$-expansion can be cast into the following forms: \begin{eqnarray} \label{corr_p_z1z2} &&G(p \to 0,z_1,z_2,L=\infty,t=0) \\ &=& {\cal G}_{\mbox{\tiny II}} \Big( \Theta(z_2-z_1) z_1^{1-\eta} \Big( \frac{z_2}{z_1} \Big)^{1-\eta_\perp} \Big( 1 + A_1 (pz_2)^{-1+\eta_\|} + \dots \Big) \nonumber \\ && + \enspace \Theta(z_1-z_2) z_2^{1-\eta} \Big( \frac{z_1}{z_2} \Big)^{1-\eta_\perp} \Big( 1 + A_1 (pz_1)^{-1+\eta_\|} + \dots \Big) \Big) , \nonumber \end{eqnarray} \begin{eqnarray} \label{corr_t_z1z2} &&G(p=0,z_1,z_2,L=\infty,t \to 0) \\ & = & {\cal G}_{\mbox{\tiny II}} \Big( \Theta(z_2-z_1) z_1^{1-\eta} \Big( \frac{z_2}{z_1} \Big)^{1-\eta_\perp} \Big( 1 + B_1 \Big( \frac{z_2}{\xi} \Big)^{-1+\eta_\|} + \dots \Big) \nonumber \\ && + \enspace \Theta(z_1-z_2) z_2^{1-\eta} \Big( \frac{z_1}{z_2} \Big)^{1-\eta_\perp} \Big( 1 + B_1 \Big( \frac{z_1}{\xi} \Big)^{-1+\eta_\|} + \dots \Big) \Big) , \nonumber \end{eqnarray} and \begin{eqnarray} \label{corr_L_z1z2} &&G(p=0,z_1,z_2,L \to \infty,t=0) \\ & = & {\cal G}_{\mbox{\tiny II}} \Big( \Theta(z_2-z_1) z_1^{1-\eta} \Big( \frac{z_2}{z_1} \Big)^{1-\eta_\perp} \Big( 1 + C_1 \Big( \frac{z_2}{L} \Big)^{-1+\eta_\|} + \dots \Big) \nonumber \\ && \enspace + \Theta(z_1-z_2) z_2^{1-\eta} \Big( \frac{z_1}{z_2} \Big)^{1-\eta_\perp} \Big( 1 + C_1 \Big( \frac{z_1}{L} \Big)^{-1+\eta_\|} + \dots \Big) \Big) . \nonumber \end{eqnarray} These expressions are valid for arbitrary $z_1$ and $z_2$ as long as the scaling variables $pz_{1,2}$, $z_{1,2}/\xi$, and $z_{1,2}/L$ are small. The explicit $\epsilon$-expansion provides the amplitudes $A_1$, $B_1$, and $C_1$ given by Eq. (\ref{A1_B1_C1}). For the special case $z_1=z_2$ Eqs. (\ref{corr_p_z1z2}) - (\ref{corr_L_z1z2}) reduce to Eqs. (\ref{corr_p}), (\ref{corr_t}), and (\ref{corr_L}). In the limits $p=0$, $t=0$, and $L=\infty$ Eqs. (\ref{corr_p_z1z2}) - (\ref{corr_L_z1z2}) reduce to Eq. (\ref{gompper_cft_z1z2}) (recall $\eta_\perp = (\eta_\|+\eta)/2$). Finally we note that Eqs. (\ref{corr_p_z1z2}), (\ref{corr_t_z1z2}), (\ref{corr_L_z1z2}), and the full film correlation function $G(p,z_1,z_2,L,t)$ up to first order in $\epsilon$ (see Eq. (\ref{g_ptL_z1z2}) in Appendix \ref{corr_for_z1z2}) satisfy the so-called product rule derived by Parry and Swain for the correlation function algebra of inhomogeneous fluids (see Eq. (2.20) in Ref. \cite{Parry:97}): \begin{eqnarray} \label{parry_product} G(p,z_1,z_2,L,t) G(p,z_2,z_3,L,t) & = & G(p,z_2,z_2,L,t) G(p,z_1,z_3,L,t) \end{eqnarray} for all $0 < z_1 \le z_2 \le z_3 < L$. The second identity derived by Parry and Swain (see Eq. (2.21) in Ref. \cite{Parry:97}) is trivially fulfilled in the disordered phase considered here because it involves the derivative of the order-parameter profile which vanishes above $T_c$. A nontrivial test of this relation would require results for the ordered phase below $T_c$. \subsection{The susceptibility} \label{excess_sus} As it has become apparent in the previous subsection, the full dependence of the correlation function $G$ on all its variables $p$, $z_1$, $z_2$, $L$, and $t$ is rather complicated. Therefore it increases the transparency to consider a spatially averaged quantity which still displays interesting specific properties of the critical behavior in a film geometry. The singular part of the total susceptibility per area defined as \begin{eqnarray} \label{def_exess_sus} \chi (L,t) & = & \int_0^L d z_1 \int_0^L d z_2 \enspace G(p=0,z_1,z_2,L,t) \end{eqnarray} provides such a reduced but still interesting quantity in that it depends only on two variables $L$ and $t$. In addition this susceptibility is directly accessible in an experiment which probes the response of a thin magnetic film on the applied external field in the limit of vanishing field strength. From the scaling properties for $G$ one obtains the following scaling property for $\chi$ (see Eqs. (\ref{G_final}) and (\ref{G_b_num})): \begin{eqnarray} \label{scaling_excess_sus} \chi (L,t) & = & B^2 (\xi_0^+)^{d+1} \Big( \frac{L}{\xi_0^+} \Big)^{3-\eta} {\cal R} f(y=L/\xi) = L^{3-\eta} {\cal G}_{\mbox{\tiny II}} f(y) \end{eqnarray} where \begin{eqnarray} \label{fy_def} f(y) & = & \int_0^1 d x_1 \int_0^1 d x_2 x_1^{1-\eta} g_{\mbox{\tiny II}} (0,x_1 y,x_2 y, x_2) \end{eqnarray} is a universal scaling function. For $y \to \infty$, i.e., $L \to \infty$ and $t$ fixed the scaling function $f(y)$ vanishes as follows: \begin{eqnarray} \label{f_large_y} f(y \to \infty) & = & {\cal A} y^{-2+\eta} + {\cal B} y^{-3+\eta} + {\cal C} y^{-3+\eta} e^{-y} + {\cal O} (e^{-2 y}) . \end{eqnarray} with \begin{eqnarray} \label{large_y_abc} {\cal A} & = & 1 - \tilde \epsilon + {\cal O} (\epsilon^2) , \nonumber \\ {\cal B} & = & -2 \Big( 1 + \tilde \epsilon \Big[ \pi \Big( \frac{1}{2} - \frac{1}{\sqrt{3}} \Big) -1 \Big] \Big) + {\cal O} (\epsilon^2) , \\ {\cal C} & = & 4 \Big( 1 - \tilde \epsilon \Big[ \frac{\pi}{2} \Big( 1 - \frac{1}{\sqrt{3}} \Big) +1 \Big] \Big) + {\cal O} (\epsilon^2) , \nonumber \end{eqnarray} so that with $\gamma = \nu (2-\eta)$ and $\gamma_s = \gamma + \nu$ \begin{eqnarray} \label{chi_large_y} \chi(L \to \infty,t) & = & B^2 (\xi_0^+)^{d+1} {\cal R} \Big\{ \frac{L}{\xi_0^+} {\cal A} t^{-\gamma} + {\cal B} t^{-\gamma_s} \Big[ 1 + \frac{{\cal C}}{{\cal B}} e^{-L/\xi} + {\cal O} (e^{-2L/\xi}) \Big] \Big\} . \end{eqnarray} The first term $(\sim t^{-\gamma})$ corresponds to the bulk contribution of the total susceptibility. (We recall that $\chi$ is the total susceptibility per area $A_\|$ of one surface and that the total volume of the system is $A_\| L$.) The universal amplitude ${\cal A}$ (Eq. (\ref{large_y_abc})) is in accordance with the corresponding known universal amplitude ratios \cite{Bervillier:76,Privman:91}. The second term $(\sim t^{-\gamma_s})$ corresponds to the sum of the excess susceptibilities of two semi-infinite systems within the surface universality class of the ordinary transition resembling the two bounding surfaces of the film. The corresponding universal amplitude ${\cal B}$ (Eq. (\ref{large_y_abc})) of the semi-infinite systems is in accordance with the corresponding result in Ref. \cite{Diehl:85:a}. Finally, the last term $\sim e^{-L/\xi}$ in Eq. (\ref{chi_large_y}) is the actual finite size contribution induced by the finite distance $L$ between the two surfaces confining the film. It is interesting to note that the structure of this finite size term ${\cal C} t^{-\gamma_s} \exp (-L/\xi)$ differs from its counterparts for the free energy and specific heat in two respects (see Eqs. (4.8) and (6.14) in Ref. [30(a)]): (i) For ordinary - ordinary boundary conditions the finite size terms of the latter two both vanish $\sim \exp (-2y)$ for large $y=L/\xi$. (ii) The prefactor ${\cal C} t^{-\gamma_s}$ is replaced by ${\cal C}' t^{-\kappa} y^{\kappa/(2\nu)}$ with $\kappa = \alpha_s -2$ (free energy) and $\kappa = \alpha_s = \alpha + \nu$ (specific heat), respectively. From the explicit result in ${\cal O}(\epsilon)$ we infer that in the case of the excess susceptibility this power law in front of the exponential is either missing or has an exponent of ${\cal O}(\epsilon^2)$. In order to render the comparison between the finite size scaling of the free energy and specific heat on one hand and of the susceptibility on the other hand more transparent we rewrite the susceptibility as \begin{eqnarray} \label{chi_SD_transparent} \chi (L,t) & = & B^2 (\xi_0^+)^{d+1} {\cal R} \Big( \frac{L}{\xi_0^+} \Big)^{\gamma_s/\nu} \Big\{ {\cal A} y^{-\gamma/\nu} + {\cal B} y^{-\gamma_s/\nu} + g(y) \Big\} \end{eqnarray} where $(2-\eta = \gamma/\nu$, $3-\eta=\gamma_s/\nu)$ \begin{eqnarray} \label{chi_SD_f_g} g(y) = f(y) - {\cal A} y^{-2+\eta} - {\cal B} y^{-3+\eta}. \end{eqnarray} The finite size scaling for the singular part ${\cal F}_{sing}$ of the free energy of a film has the similar form $(d=(2-\alpha)/\nu$, $d-1=(2-\alpha_s)/\nu)$ (see Eq. (4.11) in Ref. [30(a)]) \begin{eqnarray} \label{krech_free_energy} \frac{{\cal F}_{sing}}{k_b T_c(\infty)} & = & \frac{A_\|}{(\xi_0^+)^{d-1}} \Big( \frac{L}{\xi_0^+} \Big)^{(\alpha_s-2)/\nu} \Big\{ A_b y^{-(\alpha-2)/\nu} + A_s y^{-(\alpha_s-2)/\nu} + \Theta(y) \Big\} . \end{eqnarray} $A_\|$ is the area of the cross section of the film. Both in Eq. (\ref{chi_SD_transparent}) and Eq. (\ref{krech_free_energy}) the first two terms correspond to the bulk and surface contribution, respectively. In both cases the curly bracket represents a universal scaling function. For the susceptibility the finite size part vanishes as \begin{eqnarray} \label{chi_vanish} g(y \to \infty) & = & {\cal C} y^{-\gamma_s/\nu} e^{-y} + {\cal O} (e^{-2 y}) \end{eqnarray} whereas for the free energy one has \begin{eqnarray} \label{krech_free_energy_vanish} \Theta (y \to \infty) & = & {\cal C}' y^{-(\alpha_s-2)/(2 \nu)} e^{-2 y} + {\cal O} (e^{-3 y}) . \end{eqnarray} At this point we note that the film susceptibility has been also discussed by Nemirovsky and Freed (see Eqs. (3.14d) and (3.16d) in Ref. \cite{Nemirovsky:86:a}). Instead of the $(p, z_1, z_2)$-representation of the propagator employed here they used a discrete spectral $(p$-$\kappa_j)$-representation. In the discrete representation the propagator for Dirichlet boundary conditions is given by \begin{eqnarray} \label{G_d_pk} G_{D,j}(p,\tau) & = & \frac{1}{p^2+\tau+\kappa_j^2}, \enspace \kappa_j = \pi (j+1)/L , \enspace j=0,1,2, \dots \enspace . \end{eqnarray} The $(p, z_1,z_2)$- and $(p$-$\kappa_j)$-representation are related by the formula \begin{eqnarray} \label{G_d_pz_and_pk} G_D(p,z_1,z_2,L,\tau) & = & \frac{2}{L} \sum_{j=0}^\infty \sin (\kappa_j z_1) \sin (\kappa_j z_2) G_{D,j}(p,\tau) . \end{eqnarray} The one-loop contribution to the total susceptibility is given by \begin{eqnarray} \label{one_loop_total_sus} &&- \frac{g}{2} \frac{n+2}{3} \int\limits_0^L d z_1 \int\limits_0^L d z_2 \int \frac{d^{d-1}q}{(2\pi)^{d-1}} \int\limits_0^L dz \\ && G_D(p=0,z_1,z,L,\tau) \enspace G_D(q,z,z,L,\tau) \enspace G_D(p=0,z,z_2,L,\tau) \nonumber \\ &=&- \frac{g}{2} \frac{n+2}{3} \int\limits_0^L d z_1 \int\limits_0^L d z_2 \int \frac{d^{d-1}q}{(2\pi)^{d-1}} \int\limits_0^L dz \left( \frac{2}{L} \right)^3 \sum_{m_1,m_2,m_3=0}^\infty \sin (\kappa_{m_1} z_1) \sin (\kappa_{m_1} z) \nonumber \\ &&G_{D,m_1}(p=0,\tau) \sin^2(\kappa_{m_2} z) G_{D,m_2}(q,\tau) \sin (\kappa_{m_3} z) \sin (\kappa_{m_3} z_2) G_{D,m_3}(p=0,\tau) \nonumber . \end{eqnarray} After performing the integrations one has to evaluate the triple sum. In their calculation of the susceptibility Nemirovsky and Freed omitted the terms $m_1 \not= m_3$ in the above sum which leads to an erroneous expression for the scaling function $f(y)$. If, however, all terms in the triple sum are properly taken into account, one obtains, as expected, the same correct result for $f(y)$ as via the $(p, z_1, z_2)$-representation. The above discussion is focused on the limit $y=L/\xi \to \infty$, i.e., on increasing the film thickness at a fixed temperature. In the opposite limit $y \to 0$ the film thickness is kept fixed and one approaches the bulk critical temperature $T_c(L=\infty)$ where $\xi$ diverges as $\xi_0^+ ((T-T_c(L=\infty))/T_c(L=\infty))^{-\nu}$. For Dirichlet boundary conditions as considered here the critical temperature of the film occurs at a lower temperature $T_c(L) < T_c(L=\infty)$. Therefore the film is not critical at $T_c(L=\infty)$ and thus the susceptibility is an analytic function of $t$ around $t= (T-T_c(L=\infty))/T_c(L=\infty) =0$. Therefore the finite size scaling function $g(y \to 0)$ has the following form: \begin{eqnarray} \label{g_for_small_y} g(y \to 0) & = & - {\cal A} y^{-\gamma/\nu} - {\cal B} y^{-\gamma_s/\nu} + {\cal D} + {\cal E} y^{1/\nu} + {\cal O} (y^{2/\nu}) \end{eqnarray} with \begin{eqnarray} \label{D_sus} {\cal D} & = & \frac{1}{12} \Big( 1 - \tilde \epsilon \Big( \frac{\pi^2}{60} + 12 b_2 +1\Big) \Big) + {\cal O} (\epsilon^2) \end{eqnarray} and \begin{eqnarray} \label{E_sus} {\cal E} & = & - \frac{1}{120} \Big( 1 + \tilde \epsilon \Big( 4 a_2 - \frac{17 \pi^2}{504} + 102 b_4 - 10 b_2 -1 \Big) \Big) + {\cal O} (\epsilon^2) \end{eqnarray} so that \begin{eqnarray} \label{f_small_y} f(y \to 0) & = & {\cal D} + {\cal E} y^{1/\nu} + {\cal O} (y^{2/\nu}) . \end{eqnarray} The numbers $a_2$, $b_2$, and $b_4$ are given in Eq. (\ref{a2b2b4}) in Appendix \ref{excess_app}. For $(n,d)=(1,3)$ the values of the amplitudes to first order in $\epsilon$ are ${\cal D} \simeq 0.08142$ and ${\cal E} \simeq -0.01375$; for ${\cal A}$ and ${\cal B}$ see Eq. (\ref{large_y_abc}). The explicit form of the scaling function and its limiting behaviors are given in Appendix \ref{excess_app}. Figure \ref{fig_f_d} (a) shows $f(y)$ within mean-field theory and within perturbation theory in first order $\epsilon$ as well as its corresponding asymptotic behaviors for large and small values of $y$, and Fig. \ref{fig_f_d} (b) displays $g(y)$ for large values of $y$. Our investigations are restricted to temperatures $T \ge T_c$. Recently Leite, Sardelich, and Coutinho-Filho (LSC) \cite{Leite:99} have analyzed amplitude ratios of the specific heat and the susceptibility above $(T>T_c)$ and below $(T<T_c)$ the bulk critical temperature in the parallel plate geometry for various boundary conditions. These amplitude ratios are functions of the scaling variable $L/\xi_\pm$ (where $\xi_\pm$ is the correlation length above $(+)$ and below $(-)$ the bulk critical temperature) and describe the surface excess and finite-size contributions of the system. Their result for the amplitude function of the susceptibility above $T_c$ (see the expression for $C_+$ in Eq. (22) in Ref. \cite{Leite:99}) can be expressed in terms of the scaling function $f(y)$ as introduced in Eq. (\ref{scaling_excess_sus}). Within this framework the results of LSC to first order in $\epsilon$ for Dirichlet boundary conditions are equivalent to the following version of the scaling function $f(y)$: \begin{eqnarray} \label{f_d_LCS} f_{LSC} (y) & = & y^{-2} \Big[ 1 - \frac{\epsilon}{3} + \frac{\epsilon}{3} \int_0^1 d s f_{1/2} \left( \sqrt{s} \frac{y}{\pi} \right) - \frac{\epsilon \pi}{6 y} \Big] + {\cal O} (\epsilon^2) \end{eqnarray} with \begin{eqnarray} \label{f12} f_{1/2} (a) & = & \int_a^\infty \frac{(u^2-a^2)^{-1/2} d u}{\exp (2 \pi u) -1} \enspace . \end{eqnarray} For small values of the scaling variable $y$ this scaling function $f_{LSC}(y)$ deviates even within mean-field theory {\em qualitatively} from the actual correct form $f(y)$ given in Eqs. (\ref{fyd_2}) and (\ref{f_small_y}). Moreover, already for $y \lesssim 10$ the difference between $f_{LSC}$ and $f$ becomes larger than 10\% in ${\cal O}(\epsilon)$ and larger then 25\% within mean-field theory. These discrepancies are due to the fact that even within mean-field theory the results of LSC do not reproduce the correct surface excess contributions \cite{Diehl:85:a} and finite-size contribution (Eq. (\ref{fyd_2})). \section{Scattering cross section} \label{crosssection} \subsection{Scattering theory} As pointed out in the Introduction the diffuse scattering of X-rays and neutrons under grazing incidence allows one to probe the local structure factor near interfaces and in thin films. In this section we discuss how the singularities of the two-point correlation function near criticality in a film, as calculated above, translate into singularities of the diffuse scattering intensity under the aforementioned experimental conditions. We consider a film $(0 \le z \le L)$ composed of a material 2 sandwiched in between two halfspaces filled with material 1 $(z < 0)$ and 3 $(z >L)$, respectively (see Fig. \ref{fig_geom}). An incoming plane wave of X-rays or neutrons with momentum ${\bf K}^i = ({\bf k}_i,q_i)$ impinges on the 1-2 interface at an angle of incidence $\alpha_i$ so that $q_i = K^i \sin \alpha$ and ${\bf k}_i = K^i \cos \alpha_i (\cos \varphi_i, \sin \varphi_i, 0)$. $\lambda = 2 \pi / K^i$ is the wavelength of the X-rays or neutrons. We assume that the media 1 and 3 are homogeneous and that the 1-2 and 2-3 interfaces are laterally flat so that their contributions to diffuse scattering can be ignored. Within the plane of incidence there is a specularly reflected wave with ${\bf K}^r = ({\bf k}_i,-q_i)$. The mean value of the electron density in the case of X-rays and of the scattering length in the case of neutrons determine the intensity of the reflected beam whereas fluctuations around the mean value give rise to scattered intensity in off-specular directions ${\bf K}^f = ({\bf k}_f, q_f < 0)$ with $q_f = -K^f \sin \alpha_f$ and ${\bf k}_f = K^f \cos \alpha_f (\cos \varphi_f, \sin \varphi_f, 0)$. We consider only elastic scattering, i.e., $K^i = K^f = K^r \equiv K$. (For the more complex case of neutron scattering under grazing incidence from magnetic systems see Ref. \cite{Dietrich:85}.) In order to proceed we assume that the mean values of the electron density or of the scattering length density in each medium is constant and varies steplike across the two interfaces 1-2 and 2-3. This gives rise to the following indices of refraction \cite{Dosch:92}: \begin{eqnarray} \label{n1n2n3} z < 0 & : & n = n_1 = 1 \\ 0 < z < L & : & n = n_2 = 1-\delta_2 + i \beta_2 \nonumber \\ z > L & : & n = n_3 = 1- \delta_3 + i \beta_3 \nonumber . \end{eqnarray} In Eq. (\ref{n1n2n3}) we consider the case that medium 1 is vacuum and the generic case for hard X-rays that $Re \enspace n < 1$ in condensed matter. Although for neutrons one can also have $Re \enspace n > 1$, in order to limit the number of possible relative values of the indices of refraction for the materials 1, 2, and 3 we do not analyze this latter case in more detail. For X-rays $\delta = \lambda^2 \frac{r_e}{2 \pi} \sum_i N_i Z_i$ and the extinction coefficient $\beta = \frac{\lambda}{4 \pi} \sum_i N_i \sigma_{a,i} \equiv \frac{\lambda \mu_{abs}}{4 \pi}$ where $r_e = \frac{e^2}{4 \pi \epsilon_0 m c^2} = 2.814 \cdot 10^{-5}$ \AA \enspace is the classical electron radius, $N_i$ the number density of atoms of species $i$ with $Z_i$ electrons and absorption cross section $\sigma_{a,i}$. For neutrons $\delta = \frac{\lambda^2}{2 \pi} \sum_i N_i b_i$ and $\beta = \frac{\lambda}{4 \pi} \sum_i N_i \sigma_{t,i}$ where $b_i$ is the nuclear scattering length of species $i$. $\sigma_{t,i}$ is the cross section taking into account incoherent scattering and nuclear reactions. Typically $\delta$ and $\beta$ are of the order $10^{-5}$. For $Re \enspace n < 1$ total external reflection occurs for $\alpha < \alpha_c$. For $L=\infty$ one has $\alpha_{c12} \simeq (2 \delta_2)^{1/2}$ whereas for $L=0$ $\alpha_{c13} \simeq (2 \delta_3)^{1/2}$. Since the angle of total reflection depends only on the difference $n(z \to -\infty) - n(z \to +\infty) > 0$, for any finite $0< L < \infty$ the incoming wave is totally reflected for $\alpha < \alpha_{c13}$, independent of the index of refraction within the film. Nonetheless the types of waves propagating in the film depend on whether $\alpha \gtrless \alpha_{c12}$ (see below). For the present setup the wave field has the form $\Psi ({\bf r}, {\bf K}^i) = e^{i {\bf k}_i \cdot {\bf r}_\|} \psi (z,\alpha)$ with \begin{eqnarray} \label{wavefield} \psi(z,\alpha) & = & \left\{ \begin{array}{lcr} e^{iq_1(\alpha)z}+r_L(\alpha)e^{-iq_1(\alpha)z}&,& z < 0 \\ s_+(\alpha) e^{iq_2(\alpha)z}+s_-(\alpha)e^{-iq_2(\alpha)z}&,& 0 \leq z \leq L \\ t_L(\alpha)e^{iq_3(\alpha)z}&,& z > L \end{array} \right. \end{eqnarray} where \begin{eqnarray} \label{ssrt} r_L(\alpha) & = & \Big( (q_1-q_2)(q_2+q_3) + e^{2iq_2L}(q_1+q_2)(q_2-q_3) \Big) / \Lambda (\alpha) , \\ s_+(\alpha) & = & 2q_1(q_2+q_3)/ \Lambda (\alpha) \nonumber , \\ s_-(\alpha) & = & 2q_1(q_2-q_3)e^{2iq_2L}/ \Lambda (\alpha) \nonumber , \\ t_L(\alpha) & = & 4q_1q_2e^{i(q_2-q_3)L}/ \Lambda (\alpha) \nonumber , \\ \Lambda (\alpha) & = & (q_1+q_2)(q_2+q_3) + e^{2iq_2L}(q_1-q_2)(q_2-q_3) \nonumber . \end{eqnarray} Since the scattering cross section is independent of the intensity of the incoming beam without loss of generality we have set the amplitude of $\Psi ({\bf r}, {\bf K}^i)$ equal to 1. The vertical components of the momentum are given by \begin{eqnarray} \label{q1q2q3} q_1 (\alpha) & = & K \sin \alpha , \\ q_j (\alpha) & = & K \sqrt{n_j^2-\cos^2 \alpha} \simeq K \sqrt{\sin^2 \alpha - 2 \delta_j + 2 i \beta_j} \nonumber \\ & = & K \sqrt{\sin^2 \alpha - \sin^2 \alpha_{c1j} + 2 i \beta_j} \enspace , \enspace j = 2,3 \nonumber . \end{eqnarray} In the limiting case that the film turns into a semi-infinite substrate, i.e., $L=\infty$ one has \begin{eqnarray} \label{si_eigenfunctions} \psi_{\infty/2}(z,\alpha) & = & \left\{ \begin{array}{lcr} e^{iq_1(\alpha)z}+r_{\infty/2}(\alpha)e^{-iq_1(\alpha)z}&, & z < 0 \\ t_{\infty/2}(\alpha)e^{iq_2(\alpha)z}&, & z \geq 0 \end{array} \right. \end{eqnarray} with \begin{eqnarray} \label{rt} r_{\infty/2}(\alpha) & = & (q_1-q_2)/(q_1+q_2) , \\ t_{\infty/2}(\alpha) & = & 2q_1/(q_1+q_2) \nonumber . \end{eqnarray} The vertical momentum components $q_j(\alpha)$ have a positive imaginary part which is due to the extinction coefficient $\beta_j$ for $\alpha > \alpha_{c1j}$ and which is present for $\alpha < \alpha_{c1j}$ even in the absence of absorption. This gives rise to an exponentially damped evanescent wave with a penetration depth $l_j = (Im \enspace q_j(\alpha))^{-1}$ which increases steeply for $\alpha \nearrow \alpha_{c1j}$ and would diverge if $\beta_j = 0$. Within the film there is a superposition of two fields $s_+(\alpha) e^{i q_2(\alpha) z}$ and $s_-(\alpha) e^{-i q_2(\alpha) z}$ (Eq. (\ref{wavefield})); in the three cases $\alpha < \alpha_{c12}$ and $\beta_2 = 0$, $\alpha > \alpha_{c12}$ and $\beta_2 \not= 0$, and $\alpha < \alpha_{c12}$ and $\beta_2 \not= 0$, $q_2(\alpha)$ has a nonzero imaginary part leading to an exponentially increasing and decreasing contribution for increasing $z$. The decreasing part corresponds to the damping of the incident wave whereas the increasing part corresponds to the damping of the reflected wave generated by the interface 2-3. Equation (\ref{wavefield}) describes the wave field $\Psi ({\bf r}, {\bf K}^i)$ in the absence of any fluctuations. This wave field is scattered at the fluctuating inhomogeneities within the film giving rise to diffuse scattering intensities in off-specular directions. The computation of this intensity requires one to specify the nature of fluctuations. In the present context this amounts to specifying the kind of system undergoing the continuous phase transition in the film and to choose the appropriate order parameter. As described in the Introduction the most promising candidates for these kind of phenomena are binary alloys undergoing a continuous order-disorder phase transition concerning the occupation of fixed lattice sites $\{ {\bf R}_l \}$. (Magnetic films are equally well suited. However, the magnetic scattering of neutrons \cite{Dietrich:85} or of X-rays is more complicated and requires separate analyses. Although the details will differ from the analysis given below, the key features of the singularities are expected to be born out similarly.) In these systems a given configuration is characterized by spin-type variables $\{ S_l = \pm 1\}$ such that $S_l = +1 (-1)$ states that the lattice site ${\bf R}_l$ is occupied by a $B (A)$ atom. Accordingly the number density of electrons for such a configuration is \begin{eqnarray} \label{number_density} \rho ({\bf r}) & = & \frac{1}{2} \sum_l \left\{ \rho_B ({\bf r} - {\bf R}_l) +\rho_A ({\bf r} - {\bf R}_l) + S_l \left[ \rho_B ({\bf r} - {\bf R}_l) -\rho_A ({\bf r} - {\bf R}_l) \right] \right\} \end{eqnarray} where $\rho_{A(B)} ({\bf r})$ is the electron number density in a single unit cell $V_{cell}$ occupied by an $A (B)$ atom. (In the case of neutron scattering $\rho ({\bf r})$ stands for the scattering length density and $\rho_{A(B)} ({\bf r}) = b_{A(B)} \delta ({\bf r})$ where $b_{A(B)}$ is the mean scattering length of the nuclei of species $A (B)$.) The ordered state of this system corresponds to a configuration in which the sign of $S_l$ alternates from one lattice site to any of the neighboring ones. In this ground state the staggered ''magnetization'' $O\!P_l = S_l e^{i}\mbox{\boldmath${^\tau}$\unboldmath}^{ _m \cdot {\bf R}_l}$ is spatially constant if the reciprocal lattice vector \boldmath$\tau$\unboldmath$_m$ of the sublattice structure is chosen such that $e^{i}\mbox{\boldmath$^\tau$\unboldmath}^{_m \cdot ({\bf R}_l -{\bf R}_{l'})} = -1$ for nearest neighbor sites ${\bf R}_l, {\bf R}_{l'}$. In the reciprocal space the positions of the reciprocal sublattice vectors \boldmath$\tau$\unboldmath$_m$ are halfway in between the reciprocal lattice vectors ${\bf G}_m$ with $e^{i {\bf G}_m \cdot {\bf R}_l}$ characterizing the underlying lattice structure of the solid. (For the sake of simplicity as far as the scattering theory is concerned we do not consider here explicitly the case of systems like $Fe_3Al$ whose description requires the introduction of several sublattices.) Upon approaching the critical temperature of the continuous order-disorder transition the thermal average $\langle O\!P_l \rangle$ vanishes qualifying $O\!P_l$ as an appropriate order parameter. In the critical contribution to the {\em bulk} scattering cross section a nonzero value of $\langle O\!P_l \rangle$ leads to superlattice Bragg peaks \cite{Dietrich:95}: \begin{eqnarray} \label{bulk_super_bragg} \left( \frac{d \sigma}{d \Omega} \right)^{Bragg}_{bulk} & = & r_e^2 \left( \frac{{\bf K}^f}{K} \times {\bf e} \right)^2 \langle O\!P_l \rangle^2 \frac{N_V}{V_{cell}} (2 \pi)^3 \sum_m \left| \tilde F e^{-W} \right|^2 \delta ({\bf K}^i-{\bf K}^f- \mbox{\boldmath$\tau$\unboldmath}_m ) \end{eqnarray} where $r_e$ is the classical electron radius, ${\bf K}^f/K$ are the directions of observation, ${\bf e}$ is the polarization vector of the incoming electromagnetic wave, $\tilde F=(F_A-F_B)/2$ where $F_{A(B)} ({\bf K}) = \int_{V_{cell}} d^3 r \rho_{A(B)} (r) e^{i {\bf K \cdot r}}$ is the atomic form factor of the atom $A(B)$, $e^{-W({\bf K})}$ is the Debye Waller factor and $N_V$ is the number of lattice sites in the sample. With the independent knowledge of all prefactors in Eq. (\ref{bulk_super_bragg}) the asymptotic temperature dependence of $\left( \frac{d \sigma}{d \Omega} \right)^{Bragg}_{bulk}$ yields $\langle O\!P_l \rangle = {\cal B}' (-t)^\beta$. As discussed in Sec. \ref{model} this experimental value for ${\cal B}'$ enters into Eq. (\ref{G_final}) and there replaces $B$ if $G(p,z_1,z_2,L,t)$ corresponds to the pair correlation function $\langle O\!P_l O\!P_{l'} \rangle$ as considered below. Similarly the singular diffuse scattering around a superlattice Bragg peak $\mbox{\boldmath$\tau$\unboldmath}_m$ is given by \begin{eqnarray} \label{bulk_diffuse} \left( \frac{d \sigma}{d \Omega} \right)^{diffuse}_{bulk} & = & r_e^2 \left( \frac{{\bf K}^f}{K} \times {\bf e} \right)^2 \left| \tilde F e^{-W} \right|^2 \sum_{{\bf R}_l, {\bf R}_{l'}} \left( \langle O\!P_l O\!P_{l'} \rangle - \langle O\!P_l \rangle \langle O\!P_{l'} \rangle \right) e^{i {\bf q} \cdot ({\bf R}_l -{\bf R}_{l'})} \\ & \rightarrow & r_e^2 \left( \frac{{\bf K}^f}{K} \times {\bf e} \right)^2 \left| \tilde F e^{-W} \right|^2 \frac{N_V}{V_{cell}} \enspace G_{bulk} ({\bf q},t) \nonumber \end{eqnarray} with ${\bf q} = {\bf K}^f -{\bf K}^{i} - \mbox{\boldmath$\tau$\unboldmath}_m$. In the second part of Eq. (\ref{bulk_diffuse}) we have performed the continuum limit replacing the lattice sums by integrals (see Eq. (\ref{op_corr_q}) and the last paragraph in Sec. \ref{model}) because for $\xi \to \infty$ the lattice structure becomes irrelevant. From studying the temperature dependence of Eq. (\ref{bulk_diffuse}) for $T > T_c$ one can infer the correlation length $\xi$ and its amplitude $\xi^+_0$ introduced in Sec. \ref{model}. We note that for $q$ small compared with the inverse lattice spacing $a$ Eqs. (\ref{op_corr_q}) and (\ref{D_B_xi_0}) can be applied to Eq. (\ref{bulk_diffuse}) provided $B$ is replaced by ${\cal B}'$ as determined from Eq. (\ref{bulk_super_bragg}). Equipped with this knowledge about the critical bulk scattering (i.e., above the angle of total reflection and for a bulk sample) we can now turn to the critical diffuse scattering from the film. Within the so-called distorted wave Born approximation and for the model of the film as described above one finds the following expression for the singular part of the coherent scattering cross section \cite{Dietrich:95}: \begin{eqnarray} \label{film_diffuse} \frac{d \sigma}{d \Omega} & = & r_e^2 \frac{A_\|}{(V^\|_{cell} a)^2} \left| \tilde F e^{-W} \right|^2 \Sigma \enspace , \\ \Sigma & = & \int\limits_0^L d z_1 \int\limits_0^L d z_2 \psi_f (z_1) \psi_i (z_1) \psi_i^\ast (z_2) \psi_f^\ast (z_2) G(p,z_1,z_2,L,t) \enspace , \nonumber \end{eqnarray} where $A_\| = N_\| V^\|_{cell}$ is the illuminated surface area where $N_\|$ is the number of lattice sites at the surface and $V^\|_{cell}$ is the two-dimensional unit cell of the surface, $a$ is the lattice spacing of the cubic lattice, $\psi_{i,f}(z) \equiv \psi(z,\alpha=\alpha_{i,f})$ (see Eq. (\ref{wavefield}) and Fig. \ref{fig_geom}), and ${\bf p} = {\bf k}^f - {\bf k}^i - \mbox{\boldmath$\tau$\unboldmath}_m$ assuming that the film surfaces are cut such that $\mbox{\boldmath$\tau$\unboldmath}_m$ is parallel to them. $G$ is the lateral Fourier transform of the two-point order parameter correlation function: \begin{eqnarray} \label{g_p_discret} G(p,z_1,z_2,L,t) & = & \frac{V^\|_{cell}}{N_\|} \sum_{{\bf r}_\|^{(m)}, {\bf r}_\|^{(m')}} e^{i {\bf p} \cdot ({\bf r}_\|^{(m)} - {\bf r}_\|^{(m')})} \\ && \times \left( \langle O\!P ({\bf r}_\|^{(m)},z_1 ) O\!P ({\bf r}_\|^{(m')},z_2 ) \rangle - \langle O\!P ({\bf r}_\|^{(m)},z_1 ) \rangle \langle O\!P ({\bf r}_\|^{(m')},z_2 ) \rangle \right) \nonumber \\ & \rightarrow & \int d^2 r_\| \enspace e^{-i {\bf p} \cdot {\bf r}_\|} \enspace G({\bf r}_\|,z_1,z_2,L,t) \nonumber \end{eqnarray} on the lattice and in the continuum limit, respectively. Thus after replacing the nonuniversal amplitude $B$ in Eq. (\ref{G_final}) by ${\cal B}'$ as obtained from Eq. (\ref{bulk_super_bragg}) for $\langle O\!P_l \rangle$ we can study the scattering cross section in Eq. (\ref{film_diffuse}) by using all the information about $G(p,z_1,z_2,L,t)$ obtained in the previous section, provided all lengths and $1/p$ are sufficiently large compared with the lattice spacing $a$ so that the continuum description is applicable. In view of the properties of the wave functions $\psi$ ($-$ only their functional forms for $0 \le z \le L$ enter into $\Sigma$ (see Eq. (\ref{wavefield}))$-$) and of the scaling form for $G(p,z_1,z_2,L,t)$ (see Eq. (\ref{G_final})) one has for $\alpha_{i,f}, \alpha_{c12,c13} \ll 1$ and $\beta_{2,3}=0$: \begin{eqnarray} \label{sigma} \Sigma & = & {\cal B}' (\xi^+_0)^{d+1} {\cal R} \left ( \frac{L}{\xi^+_0} \right)^{3-\eta} \sigma \big( p \xi, \frac{L}{\xi}, \frac{l_i}{L}, \frac{l_f}{L}, \frac{\alpha_i}{\alpha_{c12}}, \frac{\alpha_{c12}}{\alpha_{c13}} \big) \end{eqnarray} where the dimensionless function $\sigma$ is given by (Eq. (\ref{G_final})) \begin{eqnarray} \label{ss_sigma} \sigma & = & \int\limits_0^1 d x_1 \int\limits_0^1 d x_2 \psi_f(z_1= x_1 L) \psi_i(z_1 = x_1 L) \psi^\ast_i (z_2 = x_2 L) \psi^\ast_f (z_2 = x_2 L) \\ && x_1^{1-\eta} g_{\mbox{\tiny II}} \big( p L x_2, \frac{L}{\xi} x_1, \frac{L}{\xi} x_2, x_2 \big) \nonumber . \end{eqnarray} The two variables $p \xi$ and $L/\xi$ of $\sigma$ stem from the scaling function of the pair correlation function whereas the dependences of $\sigma$ on $l_i/L$, $l_f/L$, $\alpha_i/\alpha_{c12}$, and $\alpha_{c12}/\alpha_{c13}$ are due to the wave functions. For $\alpha_{i,f} < \alpha_{c12}$ \begin{eqnarray} \label{lilf} l_{i,f} & = & \frac{l_0^{(2)}}{\sqrt{1- \left( \frac{\alpha_{i,f}}{\alpha_{c12}} \right)^2}} \end{eqnarray} correspond to the {\em penetration depths of} the incoming $(i)$ and outgoing $(f)$ evanescent wave, respectively, {\em within the film material 2}. $l_0^{(2)} = (K \alpha_{c12})^{-1}$ is the minimal penetration depth $l_{i,f} (\alpha_{i,f}=0)$ in the film material. Typically $l_0$ is of the order of 30 \AA \cite{Dosch:92}. For $\alpha_i > \alpha_{c12}$ and $\alpha_f > \alpha_{c12}$ the corresponding quantities $l_i$ and $l_f$, respectively, are purely imaginary. \subsection{Interplay of length scales} \label{interplay} The scattering cross section reflects the rich interplay of five length scales: $1/p$, $\xi$, $l_i$, $l_f$, and $L$. Scaling reduces that to four independent scaling variables; moreover there is a parametric dependence on $\alpha_i/\alpha_{c12}$ and on the material constant $\alpha_{c12}/\alpha_{c13}$. It is beyond the scope of the present analysis to provide an exhaustive discussion of the full dependence on all these variables. Instead we discuss some general aspects and analyze a few specific cases in more detail in order to highlight the key features of the diffuse scattering intensity. The following cases have to be distinguished (for $T \ge T_c$): \begin{itemize} \item[I a)] $l_{i,f} \ll L$ and total reflection at 1-3 interface: $\frac{d \sigma}{d \Omega}$ is proportional to the scattering volume $A_\| min (l_i,l_f)$ \begin{itemize} \item[1.] $\xi \ll l_{i,f} \ll L$: bulk behavior convoluted with evanescent waves \item[2.] $\xi \sim l_{i,f} \ll L$: crossover bulk / $\frac{\infty}{2}$ surface behavior convoluted with evanescent waves \item[3.] $l_{i,f} \ll \xi \ll L$: $\frac{\infty}{2}$ surface behavior convoluted with evanescent waves \item[4.] $l_{i,f} \ll \xi \sim L$: $\frac{\infty}{2}$ surface behavior plus distant wall correction convoluted with evanescent waves \item[5.] $\l_{i,f} \ll L \ll \xi$: film behavior near one wall convoluted with evanescent waves \end{itemize} \item[I b)] $l_{i,f} \ll L$ and no total reflection at 1-3 interface: the difference to I a) is exponentially small, i.e., $\sim e^{-LK}$. (The volume contribution to $\frac{d \sigma}{d \Omega}$ from material 3 is insignificant because it does not exhibit critical fluctuations.) \item[II a)] $l_{i,f} \sim L$ and total reflection at 1-3 interface: crossover between $\frac{d \sigma}{d \Omega} \sim A_\| min(l_i,l_f)$ to $\frac{d \sigma}{d \Omega} \sim A_\| L$ \begin{itemize} \item[1.] $\xi \ll l_{i,f} \sim L$: $\frac{\infty}{2}$ surface behavior convoluted with film wave functions \item[2.] $\xi \sim l_{i,f} \sim L$: crossover bulk / $\frac{\infty}{2}$ surface behavior convoluted with film wave functions \item[3.] $l_{i,f} \sim L \ll \xi \to \infty$: film behavior convoluted with film wave functions \end{itemize} \item[II b)] $l_{i,f} \sim L$ and no total reflection at 1-3 interface: crossover $\frac{d \sigma}{d \Omega} \sim A_\| min(l_i,l_f) \to A_\| L$ (Again, the volume contribution from material 3 is regarded to be insignificant and is not taken into account.) \item[III a)] $l_{i,f} \gg L$ and total reflection at 1-3 interface: $\frac{d \sigma}{d \Omega} \sim A_\| L$ \begin{itemize} \item[1.] $\xi \ll L \ll l_{i,f}$: bulk behavior convoluted with film wave functions \item[2.] $\xi \sim L \ll l_{i,f}$: crossover between bulk and film behavior (including two surface contributions and distant wall corrections) convoluted with film wave functions \item[3.] $L \ll \xi \ll l_{i,f}$: film behavior convoluted with film wave functions \item[4.] $L \ll \xi \sim l_{i,f}$: film behavior convoluted with film wave functions \item[5.] $L \ll l_{i,f} \ll \xi \to \infty$: film behavior convoluted with film wave functions \end{itemize} \item[III b)] $l_{i,f} \gg L$ and no total reflection at 1-3 interface: $\frac{d \sigma}{d \Omega} \sim A_\| L$ (The volume contribution from material 3 is regarded to be insignificant.) \item[IV a)] $l_{i,f}$ imaginary and total reflection at 1-3 interface: $\frac{d \sigma}{d \Omega} \sim A_\| L$ \begin{itemize} \item[1.] $\xi \ll L$: three-dimensional bulk behavior probed by undistorted plane waves \item[2.] $L \ll \xi$: film behavior probed by undistorted plane waves \end{itemize} \item[IV b)] $l_{i,f}$ imaginary and no total reflection at 1-3 interface: $\frac{d \sigma}{d \Omega} \sim A_\| L$ (in addition to an insignificant volume contribution from material 3). \end{itemize} \subsection{Susceptibility from the scattering cross section} \label{sus_cross} For large penetration depths $l_{i,f} \gg L$ the product of wave fields in Eq. (\ref{ss_sigma}) is approximately constant. In this case for $p=0$ the universal scaling function $\sigma$ of Eq. (\ref{ss_sigma}) reduces up to a prefactor to the scaling function $f$ of the total susceptibility (see Eqs. (\ref{scaling_excess_sus}) and (\ref{fy_def})), i.e., \begin{eqnarray} \label{cross_to_suz} \sigma \left( \frac{L}{\xi} \right) = \sigma \left( p\xi=0, \frac{L}{\xi} , \frac{l_i}{L}=\infty, \frac{l_f}{L}=\infty, \frac{\alpha_i}{\alpha_{c12}} < 1, \frac{\alpha_{c12}}{\alpha_{c13}} \right) & \sim & f \left( \frac{L}{\xi} \right) . \end{eqnarray} In this limit the dependences on $\alpha_i/\alpha_{c12}$ and on $\alpha_{c12}/\alpha_{c13}$ drop out for $\alpha_i < \alpha_{c13}$; for $\alpha_i > \alpha_{c13}$ there is an insignificant bulk contribution from material 3. The five different cases 1. $-$ 5. in III a) are characterized by the various contributions of asymptotic behaviors to the scaling function $\sigma(y=L/\xi) \sim f(y)$ (see Eqs. (\ref{f_large_y}) and (\ref{f_small_y})), i.e., bulk: ${\cal A}y^{-2+\eta}$, surface: ${\cal B}y^{-3+\eta}$, distant wall: ${\cal C}y^{-3+\eta}e^{-y}$, and film behavior: ${\cal D}+{\cal E}y^{1/\nu}$. In Fig. \ref{fig_sigma_suz} we show the normalized scaling function of the scattering cross section $\sigma_0(y) = \sigma (y) / \sigma (0)$ (Eq. (\ref{ss_sigma})) within mean-field and within first-order perturbation theory as well as the asymptotic behaviors of the normalized scaling function $f_0(y) = f(y)/f(0)$ of the total susceptibility $f(y \to 0)$ (Eq. (\ref{f_small_y})) and $f(y \to \infty)$ (Eq. (\ref{f_large_y})) using mean-field exponents and amplitudes and best values for the exponents and amplitudes to first order in $\epsilon$, respectively. The cases III a) or b) with lateral momentum $p=0$ are the appropriate scattering setups in order to measure the various asymptotic behaviors of the total susceptibility by varying the temperature. Figure \ref{fig_f_durch_sigma} shows the ratio of the normalized scaling functions $f_0(y)/\sigma_0(y) = \frac{f(y)}{\sigma (y)} \frac{\sigma(0)}{f(0)}$ for all four cases I a) $-$ IV a) within mean-field theory (MFT) and within perturbation theory (PT), respectively. For large penetration depths $l_{i,f} \gg L$ (see case III a)) the deviation of the scaling function $\sigma_0 (y)$ of the scattering cross section from the scaling function $f_0(y)$ of the total susceptibility is small (see the solid lines in Fig. \ref{fig_f_durch_sigma} (a) and (b)). If the penetration depths are of the order of the film thickness, $l_{i,f} \sim L$ (see case II a)), the wavefields in Eq. (\ref{ss_sigma}) contribute and the deviation from the total susceptibility becomes visible at large values of the scaling variable $y$. For $y \to 0$ and $y \to \infty$ the dotted lines attain constant values so that there are the same critical exponents but different amplitudes for the leading asymptotic behaviors of $\sigma_0$ and $f_0$. If the penetration depths are smaller than the film thickness, $l_{i,f} \ll L$ (see case I a)), this deviation is much more pronounced (see dashed lines). The difference in the amplitudes is decreased if $\alpha_{i,f} > \alpha_{c12}$, i.e., for imaginary $l_{i,f}$ (see case IV a) and dashed-dotted lines). \subsection{Dependence on the film thickness} \label{L_dependence} In order to reveal the $(1/L)^{-1+\eta_\|}$ cusp singularity in the scattering cross section. We consider the case $p=t=0$ and introduce the corresponding scattering function \begin{eqnarray} \label{sigma_film_L} \Sigma_L & = & \int\limits_0^L d z_1 \int\limits_0^L d z_2 \psi_f (z_1) \psi_i (z_1) \psi_i^\ast (z_2) \psi_f^\ast (z_2) G(p=0,z_1,z_2,L,t=0), \end{eqnarray} where the wave fields are given in Eq. (\ref{wavefield}). For the correlation function $G$ we use the asymptotic expansion given by Eq. (\ref{corr_L_z1z2}). Furthermore we introduce the scattering function of the semi-infinite system \begin{eqnarray} \label{sigma_si_L} \Sigma_\frac{\infty}{2} & = & \int\limits_0^\infty d z_1 \int\limits_0^\infty d z_2 \psi^{(f)}_\frac{\infty}{2} (z_1) \psi^{(i)}_\frac{\infty}{2} (z_1) \psi^{(i)\ast}_\frac{\infty}{2} (z_2) \psi^{(f)\ast}_\frac{\infty}{2} (z_2) G(p=0,z_1,z_2,L=\infty,t=0), \end{eqnarray} with the wave fields and the correlation function given in Eq. (\ref{si_eigenfunctions}) and Eq. (\ref{gompper_firstorder_z1z2}), respectively. The ratio of Eqs. (\ref{sigma_film_L}) and (\ref{sigma_si_L}) defines the scattering function \begin{eqnarray} \label{scattering_L} S(LK;\alpha_i,\alpha_f,\alpha_{c12},\alpha_{c13},\beta_2,\beta_3) & = & \frac{\Sigma_{L}}{\Sigma_\frac{\infty}{2}} \end{eqnarray} for $p=t=0$, where the film thickness $L$ and the momentum $K$ of the scattered wave form the scaling variable, the angles $\alpha=\{\alpha_{i,f},\alpha_{c12,c13}\}$ characterize the scattering geometry, and the extinction coefficients $\beta=\{\beta_2, \beta_3\}$ take into account photo absorption. From Eqs. (\ref{dd_def}) - (\ref{aa_si}) in Appendix \ref{details_crosssection} one obtains the asymptotic expansion \begin{eqnarray} \label{s_LK_asymptotic} S(LK \to \infty;\alpha,\beta) & = & s_0 (LK;\alpha,\beta) + s_1(LK;\alpha,\beta) C_1 \Big( \frac{1}{LK} \Big)^{-1+\eta_\|} + \dots , \end{eqnarray} with \begin{eqnarray} \label{s0_s1_LK} s_0 (LK \to \infty;\alpha,\beta) & \sim & 1 + s_0^{(1)}(LK;\alpha,\beta) e^{-LK s_0^{(2)}(\alpha,\beta)}, \\ s_1 (LK \to \infty;\alpha,\beta) & \sim & s_1^{(0)}(LK=\infty;\alpha,\beta) + s_1^{(1)}(LK;\alpha,\beta) e^{-LK s_1^{(2)}(\alpha,\beta)} \nonumber , \end{eqnarray} and $C_1$ given by Eq. (\ref{A1_B1_C1}). The functions $s_0$ and $s_1$ carry the $L$-dependence of the wave functions (see Appendix \ref{ssss}). The $L$-dependence due to the correlation function is given by the cusp singularity $C_1 (1/LK)^{-1+\eta_\|}$. The range of the values of the scaling variable $(LK)^{-1}$ is limited by the validity of the continuum theory applied here, i.e., $L \gtrsim 30$ \AA \enspace and the distorted-wave Born approximation, i.e., $K \gtrsim 1$ \AA $^{-1}$, leading to $(LK)^{-1} \lesssim \frac{1}{30}$. For small angles, i.e., for grazing incidence scattering experiments Eq. (\ref{q1q2q3}) reduces \begin{eqnarray} \label{q1q2q3_small} q_1 (\alpha) & \simeq & K \alpha , \\ q_j (\alpha) & \simeq & K \sqrt{\alpha^2 - \alpha_{c1j}^2 + 2 i \beta_j} , \enspace j = 2,3 \nonumber . \end{eqnarray} Photo absorption, $\beta_2 \not=0$, or evanescent scattering, $\alpha_{i,f} < \alpha_{c12}$, turn $q_2$ into an imaginary quantity, which leads to a real part of $s_0^{(2)}$ and $s_1^{(2)}$ in Eq. (\ref{s0_s1_LK}). If at least one angle $\alpha_i$ or $\alpha_f$ is larger than the critical angle $\alpha_{c12}$ the functions $s_0^{(1)}$ and $s_1^{(2)}$ have a real and an imaginary part. In the latter case one expects that the scattering function $S$ in Eq. (\ref{s_LK_asymptotic}) exhibits an oscillatory behavior. In Fig. \ref{fig_s_LK} we show the exponentiated scattering function and its asymptotic form for various scattering geometries. The exponentiated form is obtained by subtracting the leading behavior of the one-loop $\epsilon$-expanded result (defined by Eq. (\ref{scattering_L})) and by adding the leading behavior (see Eq. (\ref{g_as})) calculated with the best available critical exponents ($\eta \simeq 0.031$, $\eta_\perp \simeq 0.75$, and $\eta_\| \simeq 1.48$). The dashed line in Fig. \ref{fig_s_LK} corresponds to the leading asymptotic behavior, if the $L$ dependence of the wave fields is neglected: \begin{eqnarray} \label{s_LK_asymptotic_stat} S(LK \to \infty;\alpha,\beta) & = & 1 + s^{(0)}_1(LK=\infty;\alpha,\beta) C_1 \Big( \frac{1}{LK} \Big)^{-1+\eta_\|} + \dots \enspace . \end{eqnarray} Thus the full lines in Fig. \ref{fig_s_LK} take into account the whole $L$ dependence stemming from both the scattering theory and the correlation function, whereas the dashed lines take into account only the leading asymptotic $L$ dependence of the correlation function. The oscillatory behavior appearing for $\alpha_i \lessgtr \alpha_{c12} \lessgtr \alpha_f$ stems from the scattering theory (see Fig. \ref{fig_s_LK}). For the case $\alpha_{i,f} < \alpha_{c12,c13}$ in Fig. \ref{fig_s_LK} half of the maximum value of the scattering function $S$ is reached for $LK \simeq 1.5 \cdot 10^{-3}$. This corresponds to a film thickness $L \simeq 600$ \AA, i.e., $200$ monolayers (with $K \simeq 1$ \AA$^{-1}$ and 1 monolayer $\simeq 3$ \AA \enspace thick); 90\% of the maximum value of $S$ is reached for $LK \simeq 5 \cdot 10^{-5}$ which corresponds to a film thickness $L \simeq 20000$ \AA \enspace or 6700 monolayers. This demonstrates the slow convergence to the semi-infinite limit. The spatial resolution is determined by the uncertainty of the film thickness. With $\Delta L \simeq 3$ \AA \enspace (1 monolayer) this gives $K \Delta L \simeq 3$ leading to a resolution of $\Delta (LK)^{-1} \sim 3/(LK)^2$, which is not visible on the scale of Fig. \ref{fig_s_LK}. Based on these considerations we conclude that the oscillations are experimentally accessible. \subsection{Emergence of cusp singularities} \label{emergence} In the following we analyze how the cusp singularities emerge in the limit of vanishing scaling variables. To his end we chose as an example a scattering function of the two scaling variables $p/K$ and $LK$. Analogous to Eq. (\ref{sigma_film_L}) we define the quantity \begin{eqnarray} \label{sigma_film_pL} \Sigma_{p,L} & = & \int\limits_0^L d z_1 \int\limits_0^L d z_2 \psi_f (z_1) \psi_i (z_1) \psi_i^\ast (z_2) \psi_f^\ast (z_2) G(p,z_1,z_2,L,t=0) . \end{eqnarray} Together with Eq. (\ref{sigma_si_L}) this leads to the scattering function \begin{eqnarray} \label{scattering_pL} S(p/K,LK;\alpha,\beta) & = & \frac{\Sigma_{p,L}}{\Sigma_\frac{\infty}{2}} \end{eqnarray} where $\alpha=\{\alpha_{i,f},\alpha_{c12,c13}\}$ denotes the set of angles and $\beta=\{\beta_{2,3}\}$ the extinction coefficients. As in Subsec. \ref{correlation_zz} and Eq. (\ref{polar_coordinates}) we introduce polar coordinates \begin{eqnarray} \label{polar_coord_scattering} \omega & = & \sqrt{(p/K)^2+(L K)^{-2}} , \enspace \varphi = \arctan(pL) \enspace , \\ (LK)^{-1} & = & \omega \cos \varphi, \enspace p/K = \omega \sin \varphi \enspace . \nonumber \end{eqnarray} This leads to the relation \begin{eqnarray} \label{polar_scattering_function} S(p/K,LK;\alpha,\beta) & = & S(\omega \sin \varphi, (\omega \cos \varphi)^{-1};\alpha,\beta) = S_{polar} (\omega, \varphi;\alpha,\beta) \end{eqnarray} so that the leading asymptotic behavior is given by \begin{eqnarray} \label{asympt_pol_scat_fctn} S_{polar} (\omega \to 0, \varphi;\alpha,\beta) & = & S_0(\varphi;\alpha,\beta) + S_1(\varphi;\alpha,\beta) \enspace \omega^{-1+\eta_\|} + \dots \end{eqnarray} with $S_0(\varphi;\alpha,\beta) = 1$. The amplitude $S_1$ of the leading asymptotic behavior $\omega^{-1+\eta_\|}$ depends not only on the polar variable $\varphi$, as it is the case for the corresponding correlation function (see Subsec. \ref{correlation_zz}), but also on the parameters $\alpha$ and $\beta$ characterizing the scattering process. Within mean-field theory this amplitude is defined in Appendix \ref{mft_crosssection} by Eq. \ref{mft_cross_pL}. In Fig. \ref{fig_s_pL} (a) we show the exponentiated scattering function $S(p/K,LK;\alpha,\beta)$ (Eq. (\ref{scattering_pL})), where we have subtracted the leading asymptotic behavior from the mean-field expression of the scattering function and added the exponentiated form. (See Eqs. (\ref{mft_cross_pLt}) and (\ref{mft_cross_pL}) in Subappendix \ref{mft_crosssection}); the scattering function $S$ (Eq. (\ref{scattering_pL})) is a sum (see Eq. (\ref{film_functions}) in Appendix \ref{ssss}) of functions of the type ${\cal S}$ as discussed in Eq. (\ref{mft_cross_pLt}) in Subappendix \ref{mft_crosssection}.) Figure \ref{fig_s_pL} (b) illustrates the emergence of the $(p/K)^{-1+\eta_\|}$ cusp for increasing film thickness, i.e., $(LK)^{-1} \to 0$. Figure Fig. \ref{fig_s_pL} (c) shows the emergence of the $(1/(LK))^{-1+\eta_\|}$ cusp for vanishing lateral momentum $p/K \to 0$. In the later case the vertical cross sections of the manifold are not monotonous; they exhibit a maximum ($\bullet$) at $1/L \not= 0$. Figure \ref{fig_s_pL} corresponds to scattering angles $\alpha_{i,f} < \alpha_{c12,c13}$ which yields a monotonous behavior of the scattering function. Analogous considerations describe the emergence of the cusp singularities in the $\xi$-$L$ and $\xi$-$p$ dependences (see Subappendix \ref{mft_crosssection}). \section{Summary} By using fieldtheoretic renormalization group theory we have studied the singular part of the two-point correlation function in a film of thickness $L$ near the critical point $T_c$ of the corresponding bulk system. For $T \ge T_c$ and Dirichlet boundary conditions we have obtained the following main results: \noindent (1) The two-point correlation function as a function of the lateral momentum ${\bf p}$ corresponding to the $d-1$ translationally invariant directions of the film geometry, the coordinates $z_1$ and $z_2$ perpendicular to the parallel surfaces, the film thickness $L$, and temperature $t=(T-T_c)/T_c$ (or equivalently the bulk correlation length $\xi = \xi_0^+ t^{-\nu}$) exhibits three cusp singularities: $p^{-1+\eta_\|}$ for $t=0$ and $L=\infty$, $(1/\xi)^{-1+\eta_\|}$ for $p=0$ and $L=\infty$, and $(1/L)^{-1+\eta_\|}$ for $p=t=0$ (see Eqs. (\ref{corr_p_z1z2}) - (\ref{corr_L_z1z2}) and Fig. \ref{fig_p_t_L}). The emergence of these three cusp singularities is revealed by studying appropriate scaling functions of two scaling variables (see Eqs. (\ref{g_pt}) - (\ref{g_tL}) and Figs. \ref{fig_H1H2H3_phi} - \ref{fig_L_cusp_pt}). (2) The film correlation function calculated up to first order perturbation theory in $\epsilon=4-d$ satisfies the so-called product rule derived by Parry and Swain for the correlation function algebra of inhomogeneous fluids in Ref. \cite{Parry:97} (see Eq. (\ref{parry_product})). (3) By setting $p=0$ and integrating over the perpendicular coordinates $z_1$ and $z_2$ we obtain the total susceptibility of the film (Eq. (\ref{def_exess_sus})). Its dependence on $L$ and $\xi$ is described by a universal scaling function $f(y=L/\xi)$ (see Eq. (\ref{scaling_excess_sus}) and Fig. \ref{fig_f_d}) and exhibits a typical film behavior: $f(y)$ is analytic for $y \to 0$ and $f(y \to \infty)$ contains the bulk-, surface-, and finite-size contributions (see Eqs. (\ref{f_small_y}) and (\ref{f_large_y}), respectively). These properties are similar to those of the specific heat of a critical film \cite{Krech:92:a}. Our results correct previous findings in the literature \cite{Nemirovsky:86:a,Leite:99} (see the discussions of Eqs. (\ref{one_loop_total_sus}) and (\ref{f_d_LCS})). (4) In view of proposed experimental tests with X-rays and neutrons under grazing incidence (see Fig. \ref{fig_geom}), as discussed in detail in Sec. \ref{secI}, we have calculated the critical diffuse scattering from the film within the so-called distorted wave Born approximation . The scattering intensity is a function of the lateral momentum transfer $p$, film thickness $L$, bulk correlation length $\xi$, penetration depths $l_{i,f}$ of the incoming $(i)$ and outgoing $(f)$ waves, the critical angles of total reflection $\alpha_{c12}$ and $\alpha_{c13}$ and the extinction coefficients $\beta_2$ and $\beta_3$ of the film (2) and of the underlying substrate (3) (see Fig. \ref{fig_geom}). (5) For various ratios of $L$, $\xi$, and $l_{i,f}$ the scattering function shows the crossover between analytic, bulk, surface, and finite-size behavior (see Figs. \ref{fig_sigma_suz} and \ref{fig_f_durch_sigma}). By varying the temperature, a scattering experiment for $p=0$ and $l_{i,f} \gg L$ gives access to the aforementioned scaling function $f(y)$ of the total susceptibility (Eq. (\ref{cross_to_suz})). (6) For $p=t=0$ the leading singular behavior of the scattering function is given by the cusp singularity $(1/LK)^{-1+\eta_\|}$, where $K$ is the momentum of the incoming wave (Eq. (\ref{s_LK_asymptotic})). The maximal scattering intensity for $L \to \infty$ is reached only very slowly. For certain scattering geometries the $L$-dependence exhibits an oscillatory behavior (see Fig. \ref{fig_s_LK}). (7) The film thickness and momentum cusp singularities of the correlation function are borne out in the scattering cross section and are analyzed in Fig. \ref{fig_s_pL}. \section*{Acknowledgements} We thank E. Eisenriegler, A. Hanke, and M. Krech for helpful discussions. This work has been supported by the German Science Foundation through Sonderforschungsbereich 237 {\em Unordnung und gro{\ss}e Fluktuationen}.
\section{Introduction} Observations of radio-loud quasars are important for investigating some of the most interesting and fundamental problems of contemporary astrophysics. The foremost of these is the investigation of causes of the ``extinction'' of luminous quasars. The space density of luminous quasars has declined by about 3 orders-of magnitude from the epoch z=2--3 to the present (e.g., Hartwick \& Schade 1989). What processes led to the such a dramatic decrease in co-moving space density? Are these processes related to their galactic and/or cluster scale environments? The answers to these questions will be important in furthering our understanding galaxy evolution. Studies of the environments of quasars can also provide insight into the AGN phenomenon in general. Of contemporary interest is the relationship between quasars and radio galaxies. A variety of schemes have been proposed to link quasars and radio galaxies through differences in their environments, viewing angle, or evolutionary state (e.g., Norman \& Miley 1984; Barthel 1989; Neff \& Hutchings 1990). In particular, the \lq\lq viewing angle\rq\rq \ scheme of Barthel (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 fuzz should be identical to those of the radio galaxies at similar radio powers and redshifts (as do some of the evolutionary unification schemes). Also, radio galaxies at high redshifts (z $\gtrsim$ 0.6) exhibit the so-called \lq\lq alignment effect\rq\rq \ (McCarthy et al. 1987; Chambers et al. 1987) namely that the radio, and the rest-frame UV and optical axes are all roughly co-linear. If indeed quasars and radio galaxies are objects differentiated only by viewing angle, then quasars might also be expected to exhibit such an alignment effect over the same redshift range as the radio galaxies. Investigating the above problems with ground-based data has been hampered by the inability of separating marginally extended ``fuzz'' from the ``blinding'' light of the quasar nucleus. But despite these difficulties, limited information about the properties of the host galaxies of quasars spanning a range of redshifts and radio properties has been obtained (see e.g., Hutchings, Crampton, \& Campbell 1984; Smith et al. 1986; Stockton \& MacKenty 1987; Veron-Cetty \& Woltjer 1990; Heckman et al. 1991). The Hubble Space Telescope is well-suited for investigating quasar host galaxies. Because of its high spatial resolution, imaging with the HST allows us to remove the contribution of the quasar nucleus and investigate the properties of the host galaxies on scales of 0.1 to a few kpc, depending on the redshift of the source. This was one of the reasons that motivated us to undertake a ``snapshot survey'' of sources in the 3CR catalogue. The ``snapshot'' mode of observing, in which gaps in the primary HST schedule are filled in with short integrations of selected targets, greatly enhances the overall efficiency of the HST and is well-suited to observing large samples of objects. The results presented here should be compared and contrasted with recent HST results by other groups. Images of low/modest redshift quasars obtained with the WFPC-2 on HST show a wide variety of host galaxy morphologies including elliptical and spiral systems as well as highly disturbed, galaxies which may be interacting with and/or accreting close companions (Bahcall et al. 1995a,b; Disney et al. 1995; Hutchings \& Morris 1995; Hooper et al. 1997). The 3CR data set allows the properties of matched samples of radio-loud quasars and radio galaxies to be compared and should ultimately allow us to investigate the properties of ``fuzz'' over a wide range of redshifts (0.3 $\lesssim$ z $\lesssim$ 2), radio luminosities (log P$_{178 MHz}$ $\sim$ 10$^{27}$ to 10$^{29.3}$ W Hz$^{-1}$) and radio types (e.g., lobe-dominated to core-dominated). Since low-frequency radio emission from radio-loud AGN is thought to be emitted isotropically, the low-frequency selection of the 3CR (178 MHz) implies that it is relatively unbiased by anisotropic emission (whether it is due to relativistic beaming, emission from an optically thick accretion disk, or due to an obscuring torus), thus the 3CR is particularly well suited for investigating the relationship between radio galaxies and quasars. Here we present and describe the image analysis and data reduction for 43 quasars from this ``3CR snapshot survey''. Other papers in this series have described the properties of the radio galaxies (de Koff et al. 1996; Baum et al. 1999; McCarthy et al. 1997) and future work will compare matched samples of radio galaxies and quasars over a wide range of redshifts. \section{Sample Selection} To select objects for the snapshot survey, we used the revised 3CR sample as defined by Bennett (1962a,b), having selection constraints of (i) flux density at 178 MHz, S(178) $>$ 9 Jy, (ii) declination $>$ 5$^\circ$, and (iii) galactic latitude, $|b| >$ 10$^\circ$. All of the sources have been optically identified and have measured redshifts (Spinrad et al. 1985; Djorgovski et al. 1988). In this paper, we report on the properties of the quasars observed during the snapshot survey of a total of 267 3CR sources. The classification of these sources as quasars are based on the compilations of Spinrad et al. (1985) and Djorgovski et al. (1988). Of the total of 53 quasars listed in Spinrad et al. (1985), we have observed 41 (3C 179 and 3C 279, which we have observed, were not listed in the Spinrad et al. quasar identifications). Inevitable scheduling constraints and conflicts with other observing programs accounts for the 12 3CR quasars that were not imaged as a part of this program. Table 1 gives a summary of the observations. \section{Characteristics of the Observed Sample} The characteristics of the observed sample of quasars are provided in Table 2 and summarized graphically in Figures 1 through 3. The redshift distribution of the observed sample is fairly flat and ranges from 0.3 to about 2.1 (Table 2 and Fig. 1) and is broadly similar to that of the entire sample of 3CR quasars, the major difference is that a few of the lowest redshift quasars are excluded. Adopting a cosmology with a Hubble constant of H$_0$ = 75 km s$^{-1}$ Mpc$^{-1}$ and a deceleration parameter of q$_0$=0.5 and using the 178 MHz radio fluxes of the sample as listed in Spinrad et al. (1985), the radio (178 MHz) power distribution of the observed sample lies in the range log P$_{178 MHz}$ = 27 --- 29.3 W Hz$^{-1}$. The distribution has a strong peak at log P$_{178 MHz}$ $\approx$ 28.5 W Hz$^{-1}$ (Figure 2 and Table 2). Hence the observed quasars are amongst the most powerful known radio sources. Within the sample, the radio emission from these sources also exhibit a wide variety of physical sizes. Again, adopting the cosmology of H$_0$ = 75 km s$^{-1}$ Mpc$^{-1}$ and q$_0$=0.5 and measuring the largest angular size of the radio sources using radio maps available in the literature (see Table 2 and \S 7), shows that these sources range from compact, galaxy-sized radio sources (largest projected linear sizes $\lesssim$10 kpc) to cluster-scale sources with sizes of a few hundred kpc (Figure 2). \section{Observations and Pipeline Reduction} The observations were taken throughout HST Cycle 4 (early 1994 to mid 1995) and the integrations times ranged from 2 to 10 minutes, with 5 or 10 minutes with being typical (see Table 1). All the quasars were observed close to the center of the Planetary Camera (PC) and imaged through the F702W filter, whose bandpass corresponds approximately to that of the Cousins R filter. The camera/filter combination was chosen to give the maximum spatial and photometric sensitivity. The large width of the F702W filter bandpass means that every quasar image has some contribution to its brightness and morphology due to one or two prominent emission lines. From the shape of the system bandpass for the F702W filter available in the HST WFPC2 Instrument Handbook (1995), the prominent lines that contribute to the emission within the images and the redshift range over which they contribute are: [OIII]$\lambda$5007 (0.19 $<$ z $<$ 0.64), H$\beta$ (0.22 $<$ z $<$ 0.67), [OII]$\lambda$3727 (0.60 $<$ z $<$ 1.20), MgII$\lambda$2798 (1.13 $<$ z $<$ 1.93). Only 3C 9 has a redshift (z=2.012) that avoids having prominent emission lines within the bandpass of the F702W filter. The images were reduced using the standard pipeline (see the HST Data Handbook, Version 2, 1995). The standard pipeline includes bias subtraction, dark count correction, flat-field correction, and a determination of the absolute sensitivity. For those objects that had two or more individual exposures (see Table 1), the separate images were combined using the STSDAS task CRREJ. This task constructs an average of the input frames and iteratively removes highly deviant pixels from the average. For those quasars with only one exposure, we used the IRAF task ``cosmicrays'' to remove the effects of cosmic rays. This task detects pixels that have significantly different value than the surrounding pixels and replaces the deviant value with the average of the surrounding pixels. Weak cosmic rays that were missed using this technique were subsequently removed by fitting the background to the area immediately surrounding the suspected cosmic ray hit and using this new value as a substitute for the old value of the counts in the affected pixel. The data were flux-calibrated using the inverse sensitivity for the F702W filter of 1.834 $\times$ 10$^{-18}$ ergs s$^{-1}$ cm$^{-2}$ \AA$^{-1}$ dn$^{-1}$ and a zero point of 22.469 (Whitmore 1995). This puts the magnitudes on the ``Vega system''. To convert to the STMAG system, which assumes a flat spectral energy distribution and has a zero point of 23.231, 0.762 magnitudes should be subtracted from the magnitudes given here. \section{Image Reduction} \subsection{PSFs and EEDs} One of the most important steps of the analysis is to quantify the amount of extended emission from these quasar images. We have attempted this using two methods. To do this we collected images of the standard stars used to calibrate the F702W filter that were near the dates of the observations. Unfortunately, there were only a few exposures (5) that were useful. These were then used in two ways. We constructed an empirical point-spread function (PSF) using these observations by adding up 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, as we discuss below, the close agreement is limited to azimuthal averages -- Tiny Tim does not reproduce the detailed two-dimensional structure of the PSF. Next, we measured the encircled energy diagrams (EEDs), defined as the fraction of the flux from a point source interior to a radius r, as a function of r. We then intercompared all the EEDs (including the PSF generated by Tiny Tim) taken through a given filter to determine the reproducibility of the EED. There was 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 but insufficient data were available to perform this comparison for individual stars/PSFs taken with the F702W filter. Fortunately, we also carried out this analysis for another of the WFPC2 filters, F555W, in the course of another HST program (Lehnert et al. 1999). For this intercomparison, we used observations of approximately 20 stars. This intercomparison implies that we can robustly detect fuzz 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 (orbital) time scale variations due to the so-called ``breathing'' of the telescope (see WFPC-2 Instrument Handbook). We have restricted ourselves to radii less than about 1.5 arc seconds due to the poorly understood large angle scattering which becomes important beyond a radius of about 2''. To quantify how much of the quasar light is extended, we compared the EEDs of each quasar to that of the PSF (both empirical and model, although it makes little difference which one is used). We accomplished this by scaling the PSF EED so that the total flux difference between the PSF EED and that of each quasar is zero inside a radius of 2 PC pixels ($\approx$0.09''). The EED analysis is conservative in that the actual underlying fuzz will certainly have a finite non-zero central surface brightness. Therefore, this analysis provides a lower limit to that amount of extended flux coming from each quasar. Moreover, the choice of using a radius of 2 pixels was not arbitrary. Experimenting with the effect of scaling by the central pixel led to a much higher variation in the structure of each stellar EED compared to the EED of the model PSF. This variation is due to our limited ability in determining the exact location of the peak of emission and also due to the fact that the central point source never falls exactly in the center of one pixel. Enlarging the area overwhich the quasar and PSF are scaled greatly reduced the variation in the structure seen in the PSF by the EED analysis. However, selecting a large area leads to very limited sensitivity to extended emission and thus a radius of 2 pixels was chosen as the most suitable choice to maintain the sensitivity to mildly extended fuzz but also minimizing systematic problems. In addition, we limited the EED analysis to a radius $\leq$30 PC pixels ($\approx$1.4''). This mimizes the problem of the large angle scattering which may not be well represented in the empirical or model PSF but may contribute to the extended emission in the quasar images. Therefore, in Table 3, we quote the amount of extended emission as a fraction of the total quasar emission within a radius of 30 PC pixels (1.4''). In Figure 4, we show all of the profiles from the EED analysis. The EED procedure assumes that the peak flux of the underlying galaxy is zero. Of course we know that this is a lower limit. To investigate how much flux we may miss in the course of such a procedure, we conducted the EED analysis for a small sample of 10 radio galaxies also observed as part of the snapshot survey. These galaxies were chosen to represent the full range of morphologies and (most importantly) central surface brightnesses seen in the data on radio galaxies (see de Koff et al. 1996; McCarthy et al. 1997). Although the images reveal non-thermal nuclei in a few objects, none are as ``PSF dominated'' as any of the quasars. Under the assumption that radio galaxies are similar to quasar hosts in morphology and central surface brightness and that the central surface brightness observed are due to the underlying galaxies and not the AGNs, we determined the EED fluxes for the radio galaxies. These fluxes were smaller than the measured total fluxes by an amounts ranging from about 8\% to 50\% (with about 30\% being the average) of the total extended light. We regard this as a reasonable underestimate of the quasar fluxes by the EED analysis. \subsection{PSF Subtraction} There are obvious limitations to using the EED analysis described above. The foremost of these is that it does not provide information on the morphology of the host galaxy. Also, it in fact only provides a lower limit on the amount of extended emission since we have assumed that the contribution to the fuzz from the central 0.03 $\sq\arcsec$ is zero. To study the host galaxy morphology, it is necessary to first remove the contribution of the quasar nuclei from the images. Therefore to provide morphological information about the host galaxies, we next attempted subtraction of a scaled PSF from each of the quasar images. We used two different PSFs, a model of the HST PSF constructed using Tiny Tim and an empirical PSF constructed by averaging several images of standard stars. We would have preferred to construct a PSF using exposures of open clusters or of outer regions of globular clusters but no such images were available from Cycle 4 observations taken through the F702W filter. Also only a limited number of exposures of standard stars were available and some of the stellar images were far (up to about 300 pixels) from the center of the PC. If there were two or more images of the same standard taken at the same position close in time, the individual exposures were averaged. Images of different standards were then summed after being aligned to the nearest pixel. The model and empirical PSFs were scaled such that their peaks were about 5 - 15\% of the highest valued pixel in the quasar image and iteratively subtracted until emission due to the diffraction of the secondary support become negligible or the flux in the central pixel of the quasar image was too small to be measured. During this procedure we noted the subtraction level where residual diffraction spikes were just above the background noise level. Also, we continued the subtraction until a negative image of the diffraction spikes appeared above the level of the noise. This procedure allowed us to estimate the uncertainty in the fraction of extended emission by observing the values where the diffraction spikes in the residual image became negative due to over-subtraction of the PSF or were still present due to under-subtraction. The relatively subjective method described above allows us to parameterize the uncertainty in the PSF subtraction procedure as a function of the total brightness of the quasar. We found that for most of the quasars, the uncertainty in the flux of the host relative to that of the total (quasar plus host) was about $\pm$7\%. For about 20\% of the sample (8 quasars), the uncertainty was larger, about $\pm$15\% of the total quasar flux. We used these estimates to categorize the quasars into three groups according to the uncertainties in the fluxes of the remnant hosts. The three categories correspond to $\sim$$\pm$0.2, $\sim$$\pm$0.4, and $\sim$$\pm$0.7 magnitudes. After conducting the subtraction process with both the model PSF and the empirical PSF for about 10 of the quasar images, we concluded that the model PSF was inadequate for PSF subtraction. There is an asymmetry in the intensity of the diffraction spikes within the PC point-spread-function. The diffraction spike along the positive direction of U3 axis (see HST WFPC2 Instrument Handbook 1995) and in a direction $-$45$^\circ$ relative to the U3 axis are more intense than those along the other two directions (the spike along the $+$U3 axis being the most intense). The model PSF does not characterize this asymmetry accurately. On the other hand, while the empirical PSF characterized this asymmetry well, because of the limited number of standard star exposures available, the empirical PSF only accurately characterized the PSF over a limited radius. Hence the empirical PSF was good for representing and removing the PSF structure from the quasars that did not have highly saturated nuclei. For the quasar images that appeared to be only mildly saturated a small residual sometimes was present along the brightest diffraction spike (the spike along the +U3 direction), $\approx$1'' from the nucleus and occupying a few pixels in diameter. This distinct residual was easily identifiable and removed using fits to the surrounding background. This correction needed to be applied to about 12 of the quasar images and those quasars are noted in Table 3. This residual removal was carried out mainly for cosmetic reasons. The residuals contained very little flux and since the residual was in every case well separated from the quasar host, it has only a small effect on the final host morphology. For four quasars (3C 205, 3C 279, 3C 351, and 3C 454.3) the nuclei were highly saturated and we do not present the results of the PSF subtraction due to their unreliability. Generally, since the exposure times for all the quasars were roughly the same, quasars with brighter nuclei were more difficult to determine reliable host morphologies and brightnesses through PSF subtraction. After PSF subtraction, the images were rotated so that north is at the top of the frame and east is to the left. Then each images was smoothed with a 4 $\times$ 4 pixel median filter to remove the effects of ``hot'' pixels and residual cosmic rays, to emphasize low surface brightness features, and to reduce the additional noise in the final image due to the noise in the image of the empirical PSF used for subtraction. Contour plots of the final images are displayed in Figure 5 and 6. We note that the relatively ``clean'' appearance of the contour plots is due to the way in which they were constructed. PSF subtracting the images leads to an increase in the overall noise of the image near the quasar. After the image was smoothed by a 4 $\times$ 4 pixel median filter we then selected the lowest contour to be at about the 3$\sigma$ noise level in the region affected by the PSF subtraction, but well away from the host galaxy. Therefore, the minimum level is relatively high compared to the noise level of the entire displayed image. Picking such a relatively high minimum contour level has the benefit of only showing morphological features that have a high certainty of being real and not artifacts of the PSF subtraction. As noted above, any additional ``cleaning'' of the images was strictly limited to the removal of the residual along the +U3 direction approximately 1'' from the nucleus in the quasars as noted in Table 3. Even this procedure had only a marginal influence on the final displayed morphology. In addition, in the one case (3C 179.0) where we subtracted individually two images taken at different times, there was close morphological agreement between the two images of the host galaxy in spite of the rather dramatic change in the total magnitude of the nucleus. \subsection{On the Differences Between the EED Analysis and PSF Subtraction} As can be seen in Table 3, there are rather large differences between the amount of resolved fluxes estimated using the EED analysis and the PSF subtraction. These differences are not surprising. First and perhaps most importantly, the EED analysis is basically an integral process and thus it is sensitive to low signal-to-noise, smoothly distributed light, whereas the PSF subtraction analysis is inherently differential and highlights the very small scale features lost in the growth-curves. Second, the EED analysis will always underestimate the amount of resolved flux since it assumes that the fraction of host light in the unresolved core (central $\sim$0.1 $\sq\arcsec$) is negligible and then scales the contribution of ``fuzz'' under this assumption. Thus one cannot simply measure the amount of flux from the central $\sim$0.1 $\sq\arcsec$ to reconcile the estimates from the EED analysis with that from the PSF analysis. Moreover, it is known that as the focus of the HST changes, different parts of the PSF are affected in different ways (C. Burrows and M. McMaster, private communication). Therefore, using the diffraction spikes to gauge when to stop subtracting scaled PSFs from the image, may not give the proper subtraction of the flux from the nuclear region. This was clearly evident for some of the quasars where the flux would almost reach zero near the nucleus as emission from the diffraction spikes disappeared into the noise. Since there is some disagreement in the amount of resolved flux between the PSF subtraction and the EED analysis, in Figure 6, we provide images of the quasars that are not classified as extended by the EED analysis, but apparently have some extended flux in the PSF subtraction analysis. There are 8 such objects. \subsection{On the Consistency of Our Results with Other Investigations} Since it is difficult to ascertain robustly whether or not our procedure for determining to what extent the quasars images are extended, it is important to compare our results with those obtained by other investigators. Of course, this is challenging given the variety of individual circumstances (HST versus ground-based data, use of adoptive optics, different filters, optical versus IR images, etc) by which quasar hosts have been observed. Limiting ourselves to comparisons with other HST programs to image quasar hosts in the optical, we can say that we find broad consistency between the results presented here and those of other programs investigating the hosts of radio loud quasars. For example, Boyce, Disney, \& Bleaken (1999) and Boyce et al. (1998), for small samples of low-z radio-loud and radio-quiet quasars, found extended to total flux ratios of roughly ten to a few tens of percent for the radio-loud quasars in their studies which imply host magnitudes consistent with our results. At moderate redshifts, 0.4 $<$ z $<$ 0.5, in a study of radio loud and radio-quiet quasar hosts, Hooper, Impey, \& Foltz (1997) again found extended to total flux ratios of roughly ten to a few tens of percent for their radio-loud subsample which also imply host magnitudes consistent with our results. At the high redshifts (z$>$1), in two small samples of radio-loud quasars, both Ridgway \& Stockton (1997; which also included radio galaxies) and Lehnert et al. (1999) find similar host galaxy magnitudes as we find here for similarly high redshift quasars. The broad agreement between the results of the study presented here and those of other studies of radio-loud quasars using the WFPC2, we are confident that our analysis is robust. However, this statement needs to be made more quantitative and we plan to make detailed comparisons between our results and those of other studies in subsequent papers. \section{The ``Alignment Effect'' for 3CR Quasars} The ``Alignment Effect'' in which the axes of the optical and radio emission roughly align is a well known effect observed in high redshift radio galaxies (Chambers et al. 1987; McCarthy et al. 1987). An interesting test of various schemes attempting to relate the properties of radio galaxies and quasars is to determine whether or not a sample of quasars also exhibits a similar alignment. To this end, we measured the position angles of the radio images from the literature as given in Table 2. These position angles were determined from the core of the radio emission along the position angle of the jet. If a jet was not obvious in the radio image used, we then measured the position angle of the highest surface brightness ``hot spot'' relative to the core. For the HST images, we measured the position angle by fitting a series of ellipses as a function of surface brightness using the STSDAS program ``ellipse'' of the PSF subtracted, rotated and median smoothed quasar images. The position angle of the optical emission was taken to be the position angle at a surface brightness of 21.5 m$_{F702W}$ arcsec$^{-2}$ for all of the quasars. This value was chosen because it was the surface brightness that was bright enough so that the ellipses that were fitted to the data gave believable results with small uncertainties for all of the quasar images (i.e., the uncertainty in the ellipticity was $<$0.07 and the uncertainty in the PA was $<$20$^\circ$). We present the measured radio and optical position angles in Tables 2 and 3 respectively and a histogram of the difference in the radio and optical position angles in Figure 7. Figure 7 shows a tendency for the difference between the position angle of the principal axes of the radio and optical emission to be less than 20$^\circ$. We present this result for two reasons. One of course is because this is an important result for schemes that attempt to unify quasars and radio galaxies based on viewing angle, evolution, or environment (e.g., Norman \& Miley 1984; Barthel 1989; Neff \& Hutchings 1990). Our main reason for presenting this result is that it lends support for our methodology for determining if the quasar image is resolved and if so, for determining the morphology of the underlying host galaxy. Of course it is important not to overstate this proposition. When comparing the ``alignment'' properties of radio galaxies and quasars one obviously needs to worry about projections (especially for the quasars), since the UV and radio are not likely to be perfectly aligned in three dimensions. In spite of this caveat, we would not expect to see any correspondence between the radio and optical if the morphology revealed through PSF subtraction happened by chance. However, we also note that there is a possible weakening of the alignment with redshift and this may be an indication of the difficulty in resolving high redshift hosts with short exposures through the WFPC2 (hence we urge some due caution in over interpreting our results). A more robust analysis of the importance of this result and a comparison of the strength of the ``alignment effect'' in quasars and radio galaxies will be presented in a subsequent paper. \section{Individual Source Descriptions} In this section, we shall describe the morphologies of the individual sources focusing on the following questions. Are there any artifacts in the images related to the PSF subtraction? What is the morphology of the extended emission (isophotal size, ellipticity, orientation)? Are there signs of interactions with nearby companions? How does the radio structure relate to the optical morphology of the host as seen in the HST data? We shall also discuss the positional offsets between the quasar nucleus and any nearby (in projection) companions seen in the contour plots and the magnitudes of these companions. \noindent {\bf 3C 9, z=2.012} 3C 9 has the highest redshift of the quasars in our sample. A short (10 minute) ground-based U-band exposure of 3C 9 in Heckman et al. (1991) did not reveal any extended structure. Our HST observations do not detect extended flux. \noindent {\bf 3C 14, z=0.469} The image of this source is approximately round. At the faintest isophotes it is preferentially extended along PA$\approx$120$^\circ$. At higher isophotal levels, the orientation of the image is close to PA$\approx$180$^\circ$. There is a diffuse galaxy about 0.7'' south and 2.6'' west of the nucleus with a total magnitude of about m$_{F702W}$ $\approx$ 23.2. The radio source has a triple morphology and is extended along PA$\approx$$-$5$^\circ$ and 170$^\circ$ (Akujor et al. 1994) --- very similar to the PA of the highest surface brightness extended optical emission. \noindent {\bf 3C 43, z=1.47} 3C 43 is optically one of the most compact objects in our sample. The PSF subtracted image has a faint extension along PAs$\approx$ 30$^\circ$ and 180$^\circ$. Emission is extended to only about 1'' down to a surface brightness of 22.9 m$_{F702W}$ arcsec$^{-2}$. 3C 43 has a compact (LAS$\sim$2.6'') and complex radio structure (Sanghera et al. 1995; Akujor et al. 1991). This complex morphology makes a direct comparison between the optical and radio emission difficult, but it is interesting to note that the ``{\bf U}-like'' structure seen in the faintest isophote to the north of the nucleus is roughly mimicked in the radio map of Sanghera et al. (1995). The most northern component of the complex radio structure has been identified as the nucleus by Spencer et al. (1991). If this is the case then the extended optical emission to the south of the optical nucleus is roughly aligned with the curved jet seen in the radio images of 3C 43 (e.g., Sanghera et al. 1995). There is a nearby (in projection) companion galaxy about 3'' north and 0.2'' east of the quasar nucleus with a total magnitude, m$_{F702W}$ $\approx$ 23.5. There is another galaxy just visible on the edge of the contour plot shown in Figure 5, which is almost certainly a foreground galaxy. \noindent {\bf 3C 47, z=0.425} 3C 47 shows signs of interaction with a small galaxy approximately 1.7 arc seconds to the northeast of the nucleus. There is a second galaxy along this same direction approximately 3.5 arc seconds from the nucleus. These galaxies have total magnitudes of 21.6 m$_{F702W}$ and 21.7 m$_{F702W}$ respectively. The elongated, off-center (i.e., not centered on the nucleus) isophotes strongly suggest that the host galaxy is interacting with one or both of the nearby (perhaps only in projection) galaxies. The 5 GHz radio map of Bridle et al. (1994) show a core, jet, and two lobe morphology. The jet is at PA$\approx$210$^\circ$ which corresponds closely with a linear feature seen in the HST image presented in Figure 5. \noindent {\bf 3C 68.1, z=1.238} The HST data are consistent with a point source and thus we do not detect any extended flux around 3C 68.1. \noindent {\bf 3C 93, z=0.357} 3C 93 possesses a host with a large angular size --- approximately 3'' in diameter (at a surface brightness of 22.9 m$_{F702W}$ arcsec$^{-2}$). The isophotes are approximately round, with the brighter isophotes being extended along the north south-direction and the faintest isophotes oriented along PA$\approx$40$^\circ$. The 1.5 and 8.4 GHz radio maps of Bogers et al. (1994) show a ``core + double lobe'' radio source with a relatively large angular size ($\sim$ 41''). The radio source is oriented along PA$\approx$45$^\circ$ and thus roughly coincident with the principal axis of the outer isophotes. The morphology of the host galaxy agrees well with the R-band image of 3C 93 presented in Hes, Barthel, and Fosbury (1996). \noindent {\bf 3C 138, z=0.759} 3C 138 is a flattened system oriented preferentially along PA$\approx$130$^\circ$. The inner isophotes become irregular with extensions along PA$\approx$ 70$^\circ$ and along PA$\approx$290$^\circ$. 3C 138 is also a so-called compact steep spectrum (CSS) radio source. The high resolution radio map of Redong et al. (1991) shows a compact source whose linear and triplet structure is extended on scales of a few tenths of an arc second along PA$\approx$70$^\circ$. This axis of emission is similar to the extension of the isophotes on scales of a few tenths of an arc second seen in the optical image. \noindent {\bf 3C 147, z=0.545} The isophotes of 3C 147 are fairly flat in the high surface brightness regions and become rounder at fainter isophotal levels. The main axis of the optical emission is at PA$\approx$55$^\circ$. The galaxy is about 2'' across down to about 22 m$_{F702W}$ arcsec$^{-2}$. 3C 147 is also a well known compact steep spectrum radio source whose size is 0.5'' and has a jet-like structure pointing out from the nucleus at PA$\approx$240$^\circ$ (van Breugel et al. 1992). Moreover, there is a blob of emission 0.4'' from the nucleus at PA$\approx$25$^\circ$. In the HST image, we also see two blobs of emission about 1.4'' to the south of 3C 147. The total magnitude of these two blobs is about 21.7 m$_{F702W}$. There is also another bright galaxy visible 3.3'' east of the nucleus. The total magnitude of this galaxy is 19.4 m$_{F702W}$. \noindent {\bf 3C 154, z=0.580} The HST data are consistent with a point source in the EED analysis. However, there is evidence for extended emission from the PSF subtraction. In Figure 6, we show the morphology of this extended emission. The F702W image of 3C 154 extended by about 1.5'' at a level down to a surface brightness of 22 m$_{F702W}$ arcsec$^{-2}$. Radio maps of 3C 154 show a classical core, double lobe morphology (e.g., Bogers et al. 1994). The source is oriented along PA$\approx$100$^\circ$ and is large (LAS$\sim$53''; Bogers et al. 1994). The position angle of the radio source corresponds to a faint extension in the HST image that reaches about 1.2'' from the nucleus down to a surface brightness of 22 m$_{F702W}$ arcsec$^{-2}$. \noindent {\bf 3C 175, z=0.768} The HST image of 3C 175 is marginally resolved. The PSF subtracted image shows a complex morphology. The inner, brighter isophotes are oriented preferentially east-west. The fainter isophotes show a ``plume'' of emission to the south-east and south of the nucleus. This ``plume'' reaches about 2'' from the nucleus (down to 22 m$_{F702W}$ arcsec$^{-2}$). The radio source also has a ``classical'' radio structure of core and two radio lobes. This triple structure is oriented along PA$\approx$240$^\circ$ (jet side). The size of the radio source is large --- LAS$\sim$56''. In the optical image there is an extension in the isophotes along PA$\approx$240$^\circ$. There is no evidence from ground-based optical images of 3C 175 that it is resolved (Malkan 1984; Hes et al. 1996). \noindent {\bf 3C 179, z=0.856} The images of 3C 179 were reduced using a different method than the rest of the sample. Two 300 second images were taken of 3C 179 separated by a period of a month. Over that time, 3C 179 increased by about 0.5 magnitudes in brightness. We therefore PSF subtracted each image individually and, aligned and rotated both residual images and then averaged the two images. The magnitude and fraction of the extended emission was obtained by comparing the average quasar brightness with that of the average host brightness. Of course the fraction of extended to total flux was different in the two images. The brightest isophotes of the fuzz are oriented preferentially in the east-west direction. There is a bright ``knot'' of emission about 0.8'' from the nucleus along a PA$\approx$270$^\circ$. The radio maps of Reid et al. (1995) show a complex morphology. The radio maps show a jet along PA$\approx$270$^\circ$ and a double lobe morphology. \noindent {\bf 3C 181, z=1.382} The image of 3C 181 does not appear to be spatially resolved in these HST data. \noindent {\bf 3C 186, z=1.063} The image of 3C 186 is not resolved according to the EED analysis. However, the PSF subtraction suggests that it might be resolved. We show the possible morphology of the host in Figure 6. The image of 3C 186 shows ``fuzz'' about 1.8'' across down to a surface brightness of 22 m$_{F702W}$ arcsec$^{-2}$. There are two significant position angles of extended emission, PA$\approx$30$^\circ$ and PA$\approx$110$^\circ$. There is a nearby (in projection) galaxy about 2.3'' from the nucleus along PA=65$^\circ$. The total magnitude of this nearby companion is m$_{F702W}$=22.2. The radio morphology is compact and is oriented along PA$\approx$140$^\circ$ (Rendong et al. 1991; Spencer et al. 1991). \noindent {\bf 3C 190, z=1.195} The PSF subtracted image of 3C 190 reveals a complex morphology. The principal axis of the optical emission is along PA$\approx$140$^\circ$. There is a clump of emission approximately 0.8'' to the east of the nucleus. The radio morphology of 3C 190 is compact (LAS $\sim$3'') and is linear having a chain of several hot spots (Spencer et al. 1991) along PA$\approx$ 240$^\circ$. In the HST image, we see a distortion in the isophotes along the PA of the radio structure. \noindent {\bf 3C 191, z=1.956} There is no evidence for a resolved component in 3C 191 according to the EED analysis, but PSF subtraction analysis suggests that it is resolved. The PSF subtracted image of 3C 191 reveals a complex morphology. The general orientation of the fuzz is along PA$\approx$0$^\circ$ and $\approx$200$^\circ$. There is a ``plume'' of emission to the southeast of the nucleus along PA$\sim$140$^\circ$. The radio morphology is a core plus double lobe morphology with a principal axis of emission along PA$\approx$165$^\circ$ (Ankujor et al. 1991). \noindent {\bf 3C 204, z=1.112} 3C 204 has a relatively high percentage of extended to total emission (almost 50\%). The host galaxy is a flat system (e$\sim$0.25) and its major axis is oriented preferentially along PA$\approx$150$^\circ$. The radio emission has a core, jet, double lobe morphology with the jet oriented along PA$\approx$275$^\circ$ (Reid et al. 1995). There appears to be a faint extension of emission along PA$\approx$275$^\circ$ in the HST image of the host galaxy. This ``finger'' of emission may be related to the radio jet seen in the radio image of Reid et al. (1995). \noindent {\bf 3C 205, z=1.534} The HST image of this quasar was saturated. No PSF subtraction or EED analysis was attempted. \noindent {\bf 3C 207, z=0.684} 3C 207 does not appear to be extended in these HST data. \noindent {\bf 3C 208, z=1.110} 3C 208 does not appear to be extended in these HST data. \noindent {\bf 3C 215, z=0.412} The optical counterpart of 3C 215 appears to have a flattened elliptical structure with its major axis oriented along PA$\approx$ 135$^\circ$. The 5 GHz radio image of Bridle et al. (1994) shows a complex structure with an inner region consisting of several high surface brightness knots of emission along an approximately east-west line, engulfed in a large, more diffuse emission oriented along PA$\approx$150$^\circ$. The radio emission is seen over a large scale, LAS$\sim$ 1'. A ground-based V-band image presented by Hes et al. (1996) shows a similar morphology to the image presented here. \noindent {\bf 3C 216, z=0.67} The HST image of 3C 216 indicates that the host galaxy is an interacting system. The faint isophotes of the quasar fuzz are not centered on the quasar nucleus, but are offset to the northeast along PA$\approx$30 --- 45$^\circ$. The nearby (in projection) galaxy is about 1.6'' to the north of the nucleus and has a magnitude of 21.9. There is another brighter galaxy just off the edge of the contour plot presented here is about 4'' to the east and 2.5'' north of the nucleus. The total magnitude of this galaxy is m$_{F702W} \approx$ 20.4. The 1.7 and 5 GHz radio images of Reid et al. (1995) show a compact radio source (LAS $\sim$ 6'') oriented along PA$\approx$40$^\circ$. Along this PA lies a radio ``hotspot'' about 1'' from the nucleus. This hotspot is approximately coincident with the ``plume'' of optical emission we see in the HST image. \noindent {\bf 3C 220.2, z=1.157} 3C 220.2 does not appear to be extended in these HST data. \noindent {\bf 3C 249.1, z=0.313} The HST image of 3C 249.1 is spectacular. The extended emission comprises about 70\% of the total light from the quasar. A narrow-band HST image centered on [OIII]$\lambda$5007 (Sparks, private comm.) shows that most of the emission to the east of the nucleus is probably [OIII] emission within the bandpass of the F702W filter (see also Stockton \& MacKenty 1987). However, the comparison with the narrow-band [OIII] image suggests that much of the light from the inner parts of the nebula is likely to be continuum emission from the host galaxy. A 5 GHz radio image of Bridle et al. (1994) reveals a core, jet, double lobe morphology oriented preferentially along PA$\approx$ 100$^\circ$. \noindent {\bf 3C 254, z=0.734} The EED analysis of 3C 254 shows no evidence for resolution, but the PSF subtraction indicates that it is probably resolved. The HST image of 3C 254 has a complex morphology. Its host galaxy is oriented approximately in the east-west direction. There are ``plumes'' of emission along PA$\approx$45$^\circ$ and 285$^\circ$. The second of these plumes corresponds to the direction of the most distant radio lobe seen in the 5 GHz radio map of Reid et al. (1995). This radio map reveals that 3C 254 has a double-lobed radio morphology with a central core. \noindent {\bf 3C 263, z=0.646} The host galaxy of 3C 263 appears to be a flat system (e$\sim$0.3), with its major axis aligned along PA$\approx$350$^\circ$. There is a nearby galaxy in projection along the major axis of the galaxy ($\approx$1.9'' from the nucleus) with a magnitude of 22.2. There is also another nearby galaxy about 0.2'' south and 1.5'' west of the nucleus. This galaxy has a total magnitude of about 22.3. The 5 GHz radio map in Bridle et al. (1994) shows a large scale (LAS$\sim$51'') core, jet, double lobe source with the jet have an orientation of $\approx$110$^\circ$. The HST image shows a ``finger'' of extended emission in the counter-jet direction (PA$\sim$300$^\circ$). \noindent {\bf 3C 268.4, z=1.400} The image of 3C 268.4 does not appear to be extended in the EED analysis. PSF subtracting the image suggests that 3C 268.4 might be extended. In Figure 6, we show the possible resolved structure of the quasar. The HST image of 3C 268.4 reveals that it is another source with a complex morphology. The long axis ($\approx$1.5'') of the host is approximately along PA = 230$^\circ$. In the faintest isophotes there is also a ``finger'' of emission pointing approximately to the south. In this direction there is a nearby (in projection) galaxy that is about 2.6'' from the nucleus and has a total magnitude of 21.1. The 1.4 and 5 GHz radio images of Reid et al. (1995) show a core, jet, double lobe source with a largest angular size of about 12''. The principal axis of the radio emission is approximately along PA = 215$^\circ$ and corresponds roughly to the principal axis of the host galaxy. \noindent {\bf 3C 270.1, z=1.519} According to the EED analysis, 3C 270.1 does not appear to be extended. However, in Figure 6, we show the morphology of the possible extended emission obtained by PSF subtraction which suggests that the image of 3C 270.1 is resolved. The host galaxy of 3C 270.1 is oriented preferentially in the east-west direction (PA=100$^\circ$). There are ``plumes'' of emission to the south and to the north-west. The high resolution radio image of Akujor et al. (1994) shows a compact (LAS$\sim$10''), ``bent'', triple source (core $+$ two lobes) along PA$\approx$180$^\circ$ and 320$^\circ$. \noindent {\bf 3C 277.1, z=0.321} The host galaxy of 3C 277.1 is large compared to those in the rest of the sample (emission is seen over 3'' down to 22 m$_{F702W}$ arcsec$^{-2}$). The overall morphology of the host is round, but distortions to the inner isophotes are seen in the directions of PA$\approx$170$^\circ$ and 310$^\circ$. The high resolution 1.7 and 5 GHz images of Reid et al. (1995) show a compact (LAS$\sim$1.5'') double oriented along PA$\approx$310$^\circ$. The position of the radio ``hotspot'' to the northwest roughly corresponds to the distortion we see in the HST image of the host galaxy. \noindent {\bf 3C 279, z=0.538} The image of this quasar was saturated. No PSF subtraction was attempted. \noindent {\bf 3C 280.1, z=1.659} The host galaxy of 3C 280.1 is compact and compared to most other hosts in the sample, its surface brightness increases rapidly with decreasing distance from the nucleus. The isophotes are round (e$\sim$0.05) and are oriented along PA$\approx$350$^\circ$. The 5 GHz radio map of Swarup, Sinha, \& Saika (1982; but see also Lonsdale et al. 1992 and Akujor et al. 1994) shows a core, two hotspots (one close to the nucleus, $\sim$ 1'' to the southeast, and a more distant one, about 12'' to the west-northwest) and then a wiggly chain of emission to the southeast along PA$\approx$120$^\circ$. This ``chain'' of radio emission is seen from about 4'' to about 11'' from the nucleus. The host galaxy, as seen in our HST image, does have an outward bending of the isophotes along the direction of and over the region of the southeastern hotspot seen in the radio maps of Swarup et al. (1982). \noindent {\bf 3C 287, z=1.055} The HST data are not spatially resolved. \noindent {\bf 3C 288.1, z=0.961} The HST data are not spatially resolved. \noindent {\bf 3C 298, z=1.436} The host galaxy of 3C 298 is dominated by two morphological features --- an ``arm'' of emission that projects from the nucleus to the south-west (PA=225$^\circ$) and then bends around to the east and a ``plume'' of emission to the north-northeast of the nucleus (PA$\approx$20$^\circ$). The radio images of Rendong et al. (1991) and van Breugel et al. (1992) show a compact triple source (LAS$\sim$1.8'') with an east-west orientation. \noindent {\bf 3C 309.1, z=0.905} The host galaxy of 3C 309.1 is a flat elliptical galaxy. The major axis of the host galaxy is oriented along PA=130$^\circ$. The high resolution radio image of Redong et al. (1991) shows 3C 309.1 to be a compact source, with a nuclear region oriented along PA$\approx$145$^\circ$ and a LAS$\sim$0.1'' and a ``lobe'' about 1'' from the nucleus along PA$\approx$95$^\circ$. The highest surface brightness isophotes of the host have an orientation roughly like that of the nuclear radio emission. \noindent {\bf 3C 334, z=0.555} The HST image of 3C 334 shows a host galaxy that is distorted, having twisted and off-center isophotes. The general orientation of the host is PA$\approx$120$^\circ$ and is 1.5'' across along its major axis (down to a surface brightness of 21 m$_{F702W}$ arcsec$^{-2}$).. The 5 GHz radio image of Bridle et al. (1994) shows a large (LAS$\sim$57'') triple (core, jet, two radio lobes) source. The radio jet emerges from the nucleus at PA$\approx$140$^\circ$ and curves around to the north. The orientation of the optical image of the host galaxy is approximately the same as that of the radio (5 GHz) image. Several ground-based images of 3C 334 show that it is extended and has a morphology similar to that presented here. For example, an [OII]$\lambda$3727 image in Hes et al. (1996) shows that 3C 334 is extended along PA$\approx$10$^\circ$ and a [OIII]$\lambda$5007 image of Lawrence (1990) shows extended line emission along PA$\approx$150$^\circ$. Both results have some morphological similarity to the image presented in this study. \noindent {\bf 3C 343, z=0.988} The image of 3C 343 is most unusual for the quasars imaged in this sample; it did not require any PSF subtraction! It appears to be a flat system with a major axis of about 2'' long oriented along PA$\approx$60$^\circ$. 3C 343 is another CSS radio source. The radio map of Rendong et al. (1991) reveals a compact (LAS $\approx$0.3'') complex radio source. There is a finger of emission (a jet?) pointing out along PA$\approx$320$^\circ$. From an analysis of an optical spectrum, Aldcroft, Bechtold, \& Elvis (1994) suggest that 3C 343 is a Seyfert 2 galaxy (i.e., the galaxy has narrow permitted and forbidden lines). Given the high radio luminosity of this galaxy, a more appropriate classification is as a radio galaxy. The characteristics of the optical spectrum from Aldcroft et al. supports our imaging data and our contention that 3C 343 does not appear to be a quasar. \noindent {\bf 3C 351.0, z=0.371} The image of this quasar was saturated. No PSF subtraction was attempted. \noindent {\bf 3C 380, z=0.692} 3C 380 appears to be marginally resolved and is perhaps an interacting system. It has two companion galaxies that appear to be immersed in common isophotes with the host galaxy of 3C 380. These galaxies are approximately 0.6'' west and 0.5'' north and 0.8'' west and 0.7'' north of the nucleus respectively. The total magnitudes of these two galaxy are m$_{F702W}$ $\approx$ 20.7 and m$_{F702W}$ $\approx$ 21.9. The radio image of van Breugel et al. (1992) at 1.4 GHz shows a very diffuse radio morphology that is strikingly similar but larger than the optical HST image shown in Figure 5. The interacting companions are engulfed in this radio emission and there is a surface brightness enhancement of the radio emission over the area of these companions (Reid et al. 1995). \noindent {\bf 3C 418, z=1.686} 3C 418 appears to have a compact host galaxy with several nearby (in projection) galaxies. The long axis of the host is only about 0.9'' across (down to a surface brightness of 22.9 m$_{F702W}$ arcsec$^{-2}$) and oriented preferentially along PA$\approx$225$^\circ$. The two nearby galaxies are 0.9''W, 0.2''N and 1.2''W, 1.2''N from the nucleus and have magnitudes of 23.8 and 23.0 m$_{F702W}$. A 15 GHz radio image of O'Dea, Barvainis, \& Challis (1988) shows a complex radio morphology with several bends in a radio ``jet'' pointing approximately along PA=330$^\circ$. This twisted jet is approximately 2'' long. The positions of the bends in the radio jet seem to correspond roughly to the positions of these nearby ``companions'' seen in the PSF subtracted HST image. However, given the magnitudes of these ``companions'', it seems very unlikely that they would be at the redshift of the quasar. \noindent {\bf 3C 432, z=1.785} The HST data for 3C 432 are not extended according to the EED analysis. However, PSF subtraction suggests that 3C 432 may in fact be resolved. The HST image of 3C 432 shows a compact host about 1.2'' in diameter down to a surface brightness of 22 m$_{F702W}$ arcsec$^{-2}$. The host is preferentially oriented along PA$\approx$ 45$^\circ$. There is a secondary ``plume'' of emission along PA$\approx$ 135$^\circ$. A 4.9 GHz radio map of Bridle et al. (1994) shows a classical core, jet, double lobed source extended over 15'' and oriented along PA$\approx$135$^\circ$. \noindent {\bf 3C 454, z=1.757} The HST image of 3C 454 reveals a round and compact ($\sim$1'' at a surface brightness of 22 m$_{F702W}$ arcsec$^{-2}$) host galaxy. The radio morphology is also compact, with a largest angular size of about 0.8'' (Rendong et al. 1991; Spencer et al. 1991). The radio emission is oriented preferentially in the north-south direction (Spencer et al. 1991). As is the case for 3C 181 and 3C 288.1, we urge caution in interpreting several of the features in the PSF subtracted image shown in Figure 5. Some of the structure seen along PA$\approx$205$^\circ$ may be due to incomplete removal of the most intense diffraction spike (i.e., the one in the +U3 direction). The questionable accuracy of the PSF removal is noted in Table 3. \noindent {\bf 3C 454.3, z=0.860} The image of this quasar was saturated. No PSF subtraction was attempted. \noindent {\bf 3C 455, z=0.5427} The image of 3C 455 is interesting. It has one of the highest extended/total brightness ratios of the entire sample ($\sim$70\%). It has a simple diffuse morphology with an embedded high surface brightness nucleus. Down to a surface brightness of about 22 mag arcsec$^{-2}$, 3C 455 is about 2'' in diameter along its major axis at PA$\approx$60$^\circ$. The radio morphology in the 8 GHz map of Bogers et al. (1994) is a core, double radio lobe type oriented along PA$\approx$245$^\circ$ (very similar to the optical major axis seen in the HST image). The spectrum of this object shown in Gelderman \& Whittle (1994) shows narrow permitted lines. Combined with our result that very little point-source subtraction was necessary, suggests that this object should be re-classified as a compact radio galaxy rather than a quasar. \section{Concluding Remarks} We can draw some general conclusions from an analysis of the images presented in Figures 5 and 6. We present HST ``snapshot'' images of a sample of 43 quasars. From a close inspection and analysis of these data we draw the following conclusions: \noindent $\bullet$ Our analysis suggests that the quasar fuzz contributes from $<$5\% to nearly 100\% in the most extreme case (about 20\% being typical) of the total light from the quasar. Although a large fraction of the objects do not appear to be resolved in the EED analysis ($\sim$40\%), in about 1/2 of those sources, the more sensitive PSF subtraction indicates the presence of a resolved component. \noindent $\bullet$ Many of the resolved sources show complex morphology with twisted, asymmetric, and/or distorted isophotes and irregular extensions. \noindent $\bullet$ In almost every case of the quasars with spatially resolved ``fuzz'', there are similarities between the radio and optical morphologies. In many cases there are features with similar radio and optical morphologies and/or the principal axes of radio and optical (continuum and line) emission are roughly aligned. This is further evidence for the reality of the structures detected in the PSF analysis. \noindent $\bullet$A significant fraction ($\sim$25\%) of sources show galaxies nearby in project (within 5'') and some ($\sim$10\% of the sources) show obvious signs of interactions with these nearby companions. These results show that the generally complex morphologies of host galaxies of quasars are influenced by the radio emitting plasma and by the presence of nearby companions. We must be cautious in interpreting these data since all of the images have a contribution from emission lines to their morphologies and brightnesses. Separation of the continuum and line contributions and constraints on the various mechanisms involved await new snapshot surveys that are being carried out of 3C sources with the linear ramp filters and another broad-band filter. A more robust and quantitative analysis of the data and the relationship between high redshift radio galaxies and quasars will be presented in a future paper. \acknowledgements This research is partially supported by HST GO grant number GO-5476.01-93A, by a program funded by the Dutch Organization for Research (NWO) and by a NATO research grant. M. D. L. would like to thank Chris Burrows, Tim Heckman, and Matt McMaster for useful suggestions. We would also like to thank an anonymous referee for her/his remarks that improved the presentation of our results. \newpage
\section{Introduction} Extended objects with $p$ spatial dimensional, known as `branes', play an essential role in revealing the non-perturbative structure of the superstring theories and $M$-theories \cite{b1,b2,b3,b4}. Antisymmetric tensor gauge fields have been widely studied in the theories of $p$-branes% \cite{a1,a2,a3}. In the context of the effective $D=10$ or $D=11$ supergravity theory a $p$-brane is a $p$-dimensional extended source for a $% (p+2)$-form gauge field strength $F.$ It is well-known that the $(p+2)$-form strength $F$ satisfies the field equation% $$ \nabla _\mu F^{\mu \mu _1\cdots \mu _{p+1}}=j^{\mu _1\cdots \mu _{p+1}} $$ where $j^{\mu _1\cdots \mu _{p+1}}$is a $(p+1)$-form tensor current and corresponding to electric source, and the dual field strength $^{*}F$ satisfies% $$ \nabla _\mu ~^{*}F^{\mu \mu _1\cdots \mu _{\tilde p+1}}=\tilde j^{\mu _1\cdots \mu _{\tilde p+1}} $$ in which $\tilde j^{\mu _1\cdots \mu _{\tilde p+1}}$ is a $(\tilde p+1)$% -form tensor current and corresponding to magnetic source \cite{strom}\cite {hull}\cite{duff}. The $\phi $-mapping theory proposed by Prof. Duan\cite{DuanGe,DuanLiu} is important in studying the topological invariant and topological structure of physics systems and has been used to study topological current of magnetic monopole \cite{DuanGe}, topological string theory \cite{DuanLiu}, topological structure of Gauss-Bonnet-Chern theorem \cite{DuanMeng2}, topological structure of the SU(2) Chern density\cite{fu} and topological structure of the London equation in superconductor \cite{DuanZhangLi}. We must pointed out that the $\phi $-mapping theory is also a powerful tools to investigate the topological defects theory \cite{DuanZhangFeng,DuanZhang,DZF}% , and here the vector field $\vec \phi $ is looked upon as the order parameters of the defects. In this paper, we present a new topological tensor current of `magnetic' $% \tilde p$-branes by making use of the $\phi $-mapping theory. One shows that the each isolated zero of the $d$-dimensional vector field $\vec \phi (x)$ corresponds to a $\tilde p$-brane$(\tilde p=D-d-1),$ and this current is proved to be the general current density of multi $\tilde p$-branes. Using this current, the generalized Nambu action for multi $\tilde p$-branes is obtained. This topological tensor current will give rise to the inner structure of the field strength $F$ including the contribution of the `magnetic' $\tilde p$-branes. Finally, we show that the charges carried by multi $\tilde p$-branes are topologically quantized and labeled by the Hopf index and Brouwer degree, the winding number of the $\phi $-mapping. \section{The topological tensor current of $\tilde p$-branes} Let $X$ be a $D$-dimensional smooth manifold with metric tensor $g_{\mu \nu } $ and local coordinates $x^\mu (\mu ,\nu =0,\cdots ,D-1)$ with $x^0=t$ as time, and let $R^d$ be an Euclidean space of dimension $d<D.$ We consider a smooth map $\phi :X\rightarrow R^d$ , which gives a $d$-dimensional smooth vector field on $X$% \begin{equation} \phi ^a=\phi ^a(x),\quad a=1,2,\cdots ,d. \end{equation} The direction unit field of $\vec \phi (x)$ can be expressed as \begin{equation} \label{unit}n^a=\frac{\phi ^a}{||\phi ||},\quad \quad ||\phi ||=\sqrt{\phi ^a\phi ^a}. \end{equation} In the $\phi $-mapping theory, to extend the theory of magnetic monopoles% \cite{DuanGe} and the topological string theory\cite{DuanLiu}, we present a new topological tensor current, with the unit `magnetic' charge $g_m$, defined as $$ \tilde j^{\mu _1\cdots \mu _{D-d}}=\frac{g_m}{A(S^{d-1})(d-1)!}(\frac 1{% \sqrt{g}})\in ^{\mu _1\cdots \mu _{D-d}\mu _{D-d+1}\mu _{D-d+2}\cdots \mu _D} $$ \begin{equation} \label{current1}\in _{a_1a_2\cdots a_d}\partial _{\mu _{(D-d+1)}}n^{a_1}\partial _{\mu _{(D-d-2)}}n^{a_2}\cdots \partial _{\mu _D}n^{a_d} \end{equation} where $g$ is the determinant of the metric tensor $g_{\mu \nu }$. Obviously, this `magnetic' tensor current is identically conserved, \begin{equation} \nabla _{\mu _1}\tilde j^{\mu _1\cdots \mu _{D-d}}=0,\quad i=1,\cdots ,D-d. \end{equation} From (\ref{unit}) we have \begin{equation} \label{ded}\partial _\mu n^a=\frac 1{||\phi ||}\partial _\mu \phi ^a+\phi ^a\partial _\mu (\frac 1{||\phi ||}) \end{equation} \begin{equation} \label{ded1}\frac \partial {\partial \phi ^a}(\frac 1{||\phi ||})=-\frac{% \phi ^a}{||\phi ||^3} \end{equation} Using the above expressions, the general tensor current can be rewritten as $$ \tilde j^{\mu _1\cdots \mu _{D-d}}=\frac{g_m}{A(S^{d-1})(d-1)!(d-2)}(\frac 1{% \sqrt{g}})\in ^{\mu _1\cdots \mu _{D-d}\mu _{D-d+1}\cdots \mu _{D\ }}\in _{\ a_1\cdots a_d} $$ \begin{equation} \partial _{\mu _{(D-d+1)}}\phi ^a\partial _{\mu _{(D-d-2)}}\phi ^{a_2}\cdots \partial _{\mu _D}\phi ^{a_d}\frac \partial {\partial \phi ^a}\frac \partial {\partial \phi ^{a_1}}(\frac 1{||\phi ||^{d-2\ }}). \end{equation} If we define a generalized Jacobians tensor as \begin{equation} \label{jaco}\in ^{a_1\cdots a_d}J^{\mu _1\cdots \mu _{D-d}}(\frac \phi x% )=\in ^{\mu _1\cdots \mu _{D-d}\mu _{D-d+1}\mu _{D-d+2}\cdots \mu _D}\partial _{\mu _{(D-d+1)}}\phi ^{a_1}\partial _{\mu _{(D-d-2)}}\phi ^{a_2}\cdots \partial _{\mu _D}\phi ^{a_d} \end{equation} and make use of the generalized Laplacian Green function relation in $\phi $% -space \begin{equation} \label{delt}\frac \partial {\partial \phi ^a}\frac \partial {\partial \phi ^a% }(\frac 1{||\phi ||^{d-2\ }})=\frac{4\pi ^{\frac d2}}{\Gamma (d/2-1)}\delta (% \vec \phi ), \end{equation} we obtain a $\delta $-function like tensor current\cite{DuanLiu} \begin{equation} \label{current}\tilde j^{\mu _1\cdots \mu _{D-d}}=g_m\delta (\vec \phi )J^{\mu _1\cdots \mu _{D-d}}(\frac \phi x)(\frac 1{\sqrt{g}}). \end{equation} We find that $\tilde j^{\mu _1\cdots \mu _{D-d}}\neq 0$ only when $\phi =0$. So, it is essential to discuss the solutions of the equations \begin{equation} \phi ^a(x)=0,\quad a=1,\cdots ,d \end{equation} Suppose that the vector field $\vec \phi (x)$ possesses $l$ isolated zeroes, according to the deduction of Ref. \cite{DuanLiu} and the implicit function theorem\cite{ggg}\cite{yang76}, when the zeroes are regular points of $\phi $% -mapping, i.e. the rank of the Jacobian matrix $[\partial _\mu \phi ^a]$ is $% d$, the solution of $\vec \phi (x)=0$ can be parameterized by \begin{equation} \label{solut}x^\mu =z_i^\mu (u^1,u^2,\cdots ,u^{D-d}),\quad i=1,\cdots ,l, \end{equation} where the subscript $i$ represents the $i$-th solution and the parameters $% u=u(u^1,\cdots ,u^{D-d})$ span a $(D-d)$-dimensional submanifold of $X$, denoted by $N_i$, which corresponds to a $\tilde p$-brane$(\tilde p=D-d-1)$ with spatial $\tilde p$-dimension and $N_i$ is its worldvolume. One see that the tensor current $\tilde j^{\mu _1\cdots \mu _{D-d}}$ is not vanished only on the worldvolume manifolds $N_i$ $(i=1,\cdots ,l),$ each of which corresponds to a $\tilde p$-brane. Therefore, every isolated zero of $\vec \phi (x)$ on $X$ corresponds to a magnetic $\tilde p$-branes. These `magnetic' $\tilde p$-branes had been formally discussed and not studied based on the topology theory \cite{nepo,teit}. Here, we must pointed out that the $\tilde p$-branes, sometimes, may be considered as topological defects \cite{duff,tp}, in this case for our theory the vector field $\phi ^a(x)$ $(a=1,\cdots ,d)$ may be looked upon as the generalized order parameters\cite{DZF} for $\tilde p$-branes. In the following, we will discuss the inner structure of the topological tensor current $\tilde j^{\mu _1\cdots \mu _{D-d}}$. It can be proved that there exists a $d$-dimensional submanifold $M$ in $X$ with the parametric equation \begin{equation} \label{trcol}x^\mu =x^\mu (v^1,\cdots ,v^d),\quad \mu =1,\cdots ,D, \end{equation} which is transversal to every $N_i$ at the point $p_i$ with \begin{equation} \label{normal}g_{\mu \nu }\frac{\partial x^\mu }{\partial u^I}\frac{\partial x^v}{\partial v^A}|_{p_i}=0,\quad I=1,\cdots ,D-d,\quad A=1,\cdots ,d. \end{equation} This is to say that the equations $\vec \phi (x)=0$ have the isolated zero points on $M$. As we have pointed in Ref. \cite{DuanMeng2,fu}, the unit vector field defined in (\ref{unit}) gives a Gauss map $n:\partial M_i\rightarrow S^{d-1}$% , and the generalized Winding Number can be given by this Gauss map $$ W_i=\frac 1{A(S^{d-1})(d-1)!}\int_{\partial M_i}n^{*}(\in _{a_1\cdots a_d}n^{a_1}dn^{a_2}\wedge \cdots \wedge dn^{a_d}) $$ $$ \qquad \qquad \qquad \quad =\frac 1{A(S^{d-1})(d-1)!}\int_{\partial M_i}\in _{a_1\cdots a_d}n^{a_1}\partial _{A_2}n^{a_2}\cdots \partial _{A_d}n^{a_d}dv^{A_2}\wedge \cdots \wedge dv^{A_d} $$ \begin{equation} \qquad \qquad \quad \qquad =\frac 1{A(S^{d-1})(d-1)!}\int_{M_i}\in ^{A_1\cdots A_d}\in _{a_1\cdots a_d}\partial _{A_1}n^{a_1}\partial _{A_2}n^{a_2}\cdots \partial _{A_d}n^{a_d}d^dv. \end{equation} where $\partial M_i$ is the boundary of the neighborhood $M_i$ of $p_i$ on $M $ with $p_i\notin \partial M_i,$ $M_i\cap M_j=\emptyset .$ Then, by duplicating the derivation of (\ref{current1}) from (\ref{current}), we obtain \begin{equation} \label{wind}W_i=\int_{M_i}\delta (\vec \phi (v))J(\frac \phi v)d^dv \end{equation} where $J(\frac \phi v)$ is the usual Jacobian determinant of $\vec \phi $ with respect to $v$% \begin{equation} \in ^{a_1\cdots a_d}J(\frac \phi v)=\in ^{A_1\cdots A_d}\partial _{A_1}n^{a_1}\partial _{A_2}n^{a_2}\cdots \partial _{A_d}n^{a_d}. \end{equation} According to the $\delta $-function theory \cite{yangd77} and the $\phi $% -mapping theory, we know that $\delta (\vec \phi (v))$ can be expanded as \begin{equation} \label{ff}\delta (\vec \phi (v))=\sum_{i=1}^l\beta _i\eta _i\delta ^d(\vec v-% \vec v(p_i)) \end{equation} on $M,$ where the positive integer $\beta _i=|W_i|$ is called the Hopf index of the map $v\rightarrow \vec \phi (v)$ and $\eta _i=sgn(J(\frac \phi v% ))|_{p_i}=\pm 1$ is the Brouwer degree\cite{DuanMeng2,DuanZhangLi}. One can find the relation between the Hopf index $\beta _i,$ the Brouwer degree $% \eta _i$, and the winding number $W_i$ \begin{equation} W_i=\beta _i\eta _i, \end{equation} One see that the Eq. (\ref{ff}) is only the expansion of $\delta (\vec \phi (x))$ on $M.$ In order to investigate the expansion of $\delta (\vec \phi (x))$ on the whole manifold $X$, we must expand the $d$-dimensional $\delta $% -function of the singular point in terms of the $\delta $-function on the singular submanifold $N_i$ which had been given in Ref. \cite{yangd77}% $$ \delta (N_i)=\int_{N_i}\delta ^D(x-z_i(u))\sqrt{g_u}d^{(D-d)}u,\quad i=1,\cdots ,l $$ in which \begin{equation} g_u=\det (g_{\mu \nu }\frac{\partial x^\mu }{\partial u^I}\frac{\partial x^\nu }{\partial u^J}),\quad I,J=1,\cdots ,(D-d). \end{equation} Then, from Eqs. (\ref{ff}), and by considering the property of the $\delta $% -function, one will obtain \begin{equation} \label{redelt}\delta (\vec \phi (x))=\sum_{i=1}^l\beta _i\eta _i\int_{N_i}\delta ^D(x-z_i(u))\sqrt{g_u}d^{(D-d)}u. \end{equation} Therefore, the general topological current of the $\tilde p$-branes can be expressed directly as \begin{equation} \label{recur}\tilde j^{\mu _1\cdots \mu _{D-d}}=(\frac 1{\sqrt{g}})J^{\mu _1\cdots \mu _{D-d}}(\frac \phi x)\sum_{i=1}^l\beta _i\eta _i\int_{N_i}\delta ^D(x-z_i(u))\sqrt{g_u}d^{(D-d)}u, \end{equation} which is a new topological current theory of $\tilde p$-branes based on the $% \phi $-mapping theory. If we define a Lagrangian as \begin{equation} L=\sqrt{\frac 1{(D-d)!}g_{\mu _1\nu _1}\cdots g_{\mu _{(D-d)}\nu _{(D-d)}}% \tilde j^{\mu _1\cdots \mu _{D-d}}\tilde j^{\nu _1\cdots \nu _{D-d}}}, \end{equation} which is just the generalization of Nielsen's Lagrangian\cite{niel}, from the above deductions, we can prove that \begin{equation} L=(\frac 1{\sqrt{g}})\delta (\vec \phi (x)). \end{equation} Then, the action takes the form \begin{equation} \label{aa}S=\int_XL\sqrt{g}d^Dx=\int_X\delta (\vec \phi (x))d^Dx. \end{equation} By substituting the formula (\ref{redelt}) into (\ref{aa}), we obtain an important result% $$ S=\int_X\sum_{i=1}^l\beta _i\eta _i\int_{N_i}\delta ^D(x-z_i(u))\sqrt{g_u}% d^{(D-d)}ud^Dx $$ \begin{equation} =\sum_{i=1}^l\beta _i\eta _i\int_{N_i}\sqrt{g_u}d^{(D-d)}u, \end{equation} i.e. \begin{equation} S=\sum_{i=1}^l\eta _iS_i, \end{equation} where $S_i=\beta _i\int_{N_i}\sqrt{g_u}d^{(D-d)}u.$ This is just the generalized Nambu action for multi $\tilde p$-branes$(\tilde p=D-d-1),$ which is the straightforward generalization of Nambu action for the string world-sheet action\cite{nambu}. Here this action for multi $\tilde p$-branes is obtained directly by $\phi $-mapping theory, and it is easy to see that this action is just Nambu action for multi-strings when $D-d=2$ \cite {DuanLiu}. \section{The gauge field corresponding to the topological current} In this section, we will study the antisymmetric tensor gauge field corresponding to the topological tensor current presented in above section. We know that $p$-branes naturally acts as the `electric' source of a rank $% p+2$ field strength \begin{equation} \label{st}F=dA, \end{equation} where $A$ is a ($p+1)$-form as the tensor gauge potential and satisfies the gauge transformation $$ A\rightarrow A+d\Lambda _p. $$ From Eq. (\ref{st}), one have the Bianchi identity \begin{equation} \label{bi}dF\equiv 0. \end{equation} And the `electric' current density associated with the source can be expressed as \begin{equation} \label{equat1}j^{\mu _1\cdots \mu _{p+1}}=\nabla _\mu F^{\mu \mu _1\cdots \mu _{p+1}}. \end{equation} Just as the usual Maxwell's equation, we know that Eqs. (\ref{st}), (\ref{bi}% ) and (\ref{equat1}) imply the presence of an `electric' charge, i.e. $p$% -branes, but no `magnetic' source \cite{duff}. Now, let us discuss the case when there exists the `magnetic' source. For this case, one must introduce another $(p+2)$-form $G$ for the magnetic source, and the field strength $F$ must be modified to \begin{equation} \label{stt}F=dA+G, \end{equation} which is the generalized field strength including the contribution of the `magnetic' source, i.e. `magnetic' branes: $\tilde p$-branes with $\tilde p% =D-p-4$. To obtain the explicit expression for $G,$ let us consider the current density corresponds to magnetic source which is given by \begin{equation} \label{equat11}\tilde j^{\mu _1\cdots \mu _{\tilde p+1}}=\nabla _\mu ~^{*}F^{\mu \mu _1\cdots \mu _{\tilde p+1}}. \end{equation} Using (\ref{stt}) and (\ref{equat11}), we obtain \begin{equation} \label{resultg}\tilde j^{\mu _1\cdots \mu _{(\tilde p+1)}}=\frac 1{\sqrt{g}}% \partial _\mu (\sqrt{g}\frac{\in ^{\mu \mu _1\cdots \mu _{\tilde p+1}\mu _{% \tilde p+2}\cdots \mu _{D-1}}}{\sqrt{g}}G_{\mu _{\tilde p+2}\cdots \mu _{D-1}}), \end{equation} It has been pointed out in the above section that the current density of the `magnetic' branes is a topological current given by Eq. (\ref{current1}), which can be rewritten as $$ \tilde j^{\mu _1\cdots \mu _{D-d}}=\frac{g_m}{A(S^{d-1})(d-1)!}(\frac 1{% \sqrt{g}})\partial _{\mu _{(D-d+1)}}(\in ^{\mu _1\cdots \mu _{D-d}\mu _{D-d+1}\mu _{D-d+2}\cdots \mu _D} $$ \begin{equation} \label{currentg}\in _{a_1a_2\cdots a_d}n^{a_1}\partial _{\mu _{(D-d-2)}}n^{a_2}\cdots \partial _{\mu _D}n^{a_d}) \end{equation} where $(D-d)=\tilde p+1,$ i.e. $\tilde p=D-d-1$. Comparing the Eq. (\ref {resultg}) to (\ref{currentg}), we can obtain \begin{equation} \label{gg}G_{\mu _1\cdots \mu _{d-1}}=\frac{(-1)^{(D-d)}g_m}{A(S^{d-1})(d-1)!% }\in _{a_1a_2\cdots a_d}n^{a_1}\partial _{\mu _1}n^{a_2}\cdots \partial _{\mu _{d-1}}n^{a_d}, \end{equation} and \begin{equation} \label{aaz}G=\frac{(-1)^{(D-d)}g_m}{A(S^{d-1})(d-1)!}\in _{a_1a_2\cdots a_d}n^{a_1}dn^{a_2}\wedge \cdots \wedge dn^{a_d}. \end{equation} Of equal interest is the `magnetic' charge carried by the multi $\tilde p$% -branes, which is given by \begin{equation} \label{gg1}Q^M=\int_\Sigma \tilde j^{\mu _1\cdots \mu _{\tilde p+1}}\sqrt{g}% d\sigma _{\mu _1\cdots \mu _{\tilde p+1}} \end{equation} where $\Sigma $ is a $d$-dimension$(d=p+3)$ hypersurface in $X,$ while $% d\sigma _{\mu _1\cdots \mu _{\tilde p+1}}$is the convariant surface element of $\Sigma $ \cite{y96}. From (\ref{equat11}) and (\ref{gg1}), it is easy to prove that% $$ Q^M=\int_{\partial \Sigma }F, $$ where $\partial \Sigma $ is the boundary of $\Sigma $ and a $(p+2)$% -dimension hypersurface. Substituting (\ref{recur}) into (\ref{gg1}), we have \begin{equation} \label{gg2}Q^M=g_m\int_\Sigma J^{\mu _1\cdots \mu _{\tilde p+1}}(\frac \phi x% )\sum_{i=1}^l\beta _i\eta _i\int_{N_i}\delta ^D(x-z_i(u))\sqrt{g_u}% d^{(D-d)}ud\sigma _{\mu _1\cdots \mu _{\tilde p+1}}, \end{equation} from (\ref{jaco}), and the relation $$ \frac 1{(\tilde p+1)!}\in ^{\mu _1\cdots \mu _{\tilde p+1}\nu _1\cdots \nu _d}d\sigma _{\mu _1\cdots \mu _{\tilde p+1}}=dx^{\nu _1}\wedge \cdots \wedge dx^{\nu _d}, $$ the expression (\ref{gg2}) can be rewritten as \begin{equation} \label{gg3}Q^M=\dot g_0\int_{\phi (\Sigma )}\sum_{i=1}^l\beta _i\eta _i\int_{N_i}\frac 1{\sqrt{g}}\delta ^D(x-z_i(u))\sqrt{g_u}% d^{(D-d)}ud^{(d)}\phi . \end{equation} Since on the singular submanifold $N_i$ we have \begin{equation} \phi ^a(x)|_{N_i}=\phi ^a(z_i^1(u),\cdots ,z_i^D(u))\equiv 0, \end{equation} which leads to \begin{equation} \partial \phi ^a\frac{\partial x^\mu }{\partial u^I}|_{N_i}=0. \end{equation} Using this expression, one can prove \begin{equation} J^{\mu _1\cdots \mu _{D-d}}(\frac \phi x)|_{\vec \phi =0}=\frac{\sqrt{g}}{% \sqrt{g_u}}\in ^{I_1\cdots I_{(D-d)}}\frac{\partial x^{\mu _1}}{\partial u^{I_1}}\cdots \frac{\partial x^{\mu _{(D-d)}}}{\partial u^{I_{(D-d)}}}. \end{equation} Then we obtain an useful formula \begin{equation} \label{import}d^{(d)}\phi \sqrt{g_u}d^{(D-d)}u=\sqrt{g}d^Dx. \end{equation} By making use of the above formula and (\ref{gg3}), we finally get \begin{equation} Q^M=g_m\sum_{i=1}^l\beta _i\eta _i\int_X\delta ^D(x-z_i(u))d^Dx=g_m\sum_{i=1}^l\beta _i\eta _i. \end{equation} The above expression shows that the $i$-th brane carries the `magnetic' charge $Q_i^M=g_m\beta _i\eta _i=g_mW_i,$ which is topologically quantized and characterized by Hopf index $\beta _i$ and Brouwer degree $\eta _i,$ the winding number $W_i$ of the $\phi $-mapping. \section{Conclusion} In this paper the $\phi $-mapping theory is introduced to study the $\tilde p $-branes theory, which is development of our former theories of magnetic monopoles and topological strings. We present a new topological tensor current of magnetic multi $\tilde p$-branes and discuss the inner structure of this current in detail. It is shown that every isolated zero of the vector field $\vec \phi $ (i.e. order parameters) is just corresponding to a magnetic brane, $\tilde p$-brane$(\tilde p=D-d-1)$. The generalized Nambu action for multi $\tilde p$-branes can be obtained directly in terms of this topological current. The topological structure of the charges carried by $% \tilde p$-branes shows that the magnetic charges is topologically quantized and labeled by the Hopf index and Brouwer degree, the winding number of the $% \phi $-mapping. The theory formulated in this paper is a new concept for topological $\tilde p$-branes based on the $\phi $-mapping theory. \section{Acknowledgment} This work was supported by the National Natural Science Foundation of China and Doctoral Science Foundation of China.
\section{Introduction} Very high energy {$\gamma$-ray}\ beams from blazars can be used to measure the intergalactic infrared radiation field, since pair-production interactions of {$\gamma$-rays}\ with intergalactic IR photons will attenuate the high-energy ends of blazar spectra \cite{sds92}. In recent years, this concept has been used successfully to place upper limits on the the intergalactic IR field (IIRF) \cite{sd93} - \cite{bil98}. Determining the (IIRF), in turn, allows us to model the evolution of the galaxies which produce it. As energy thresholds are lowered in both existing and planned ground-based air Cherenkov light detectors \cite{knp}, cutoffs in the {$\gamma$-ray}\ spectra of more distant blazars are expected, owing to extinction by the IIRF. These can be used to explore the redshift dependence of the IIRF \cite{ss97}, \cite{ss98}. There are now 66 ``grazars'' ($\gamma$-ray blazars) which have been detected by the {\it EGRET} team \cite{3egret}. These sources, optically violent variable quasars and BL Lac objects, have been detected out to a redshift greater that 2. Of all of the blazars detected by {\it EGRET}, only the low-redshift BL Lac, Mrk 421 ($z = 0.031$), has been seen by the Whipple telescope \cite{punch92}. The fact that the Whipple team did not detect the much brighter {\it EGRET} source, 3C279, at TeV energies \cite{vac90}, \cite{ker93} is consistent with the predictions of a cutoff for a source at its much higher redshift of 0.54 \cite{sds92}. So too are the further detections of three other close BL Lacs ($z < 0.12$), {\it viz.}, Mrk 501 ($z = 0.034$) \cite{quinn96}, 1ES2344+514 ($z = 0.044$)\cite{cat98}, and PKS 2155-304 ($z = 0.117$) \cite{chad98} which were too faint at GeV energies to be seen by {\it EGRET}\footnotemark\footnotetext{PKS 2155-304 was seen in one observing period by {\it EGRET} as reported in the Third EGRET Catalogue \cite{3egret}}. The formulae relevant to absorption calculations involving pair-production are given and discussed in Ref. \cite{sds92}. For $\gamma$-rays in the TeV energy range, the pair-production cross section is maximized when the soft photon energy is in the infrared range: \begin{equation} \lambda (E_{\gamma}) \simeq \lambda_{e}{E_{\gamma}\over{2m_{e}c^{2}}} = 2.4E_{\gamma,TeV} \; \; \mu m \end{equation} where $\lambda_{e} = h/(m_{e}c)$ is the Compton wavelength of the electron. For a 1 TeV $\gamma$-ray, this corresponds to a soft photon having a wavelength near the K-band (2.2$\mu$m). (Pair-production interactions actually take place with photons over a range of wavelengths around the optimal value as determined by the energy dependence of the cross section; see eq. (3).) If the emission spectrum of an extragalactic source extends beyond 20 TeV, then the extragalactic infrared field should cut off the {\it observed} spectrum between $\sim 20$ GeV and $\sim 20$ TeV, depending on the redshift of the source \cite{ss97}, \cite{ss98}. \section{Absorption of Gamma-Rays at Low Redshifts} Stecker and De Jager \cite{sd98} (hereafter SD98) have recalculated the absorption coefficient of intergalactic space using a new, empirically based calculation of the spectral energy distribution (SED) of intergalactic low energy photons by Malkan and Stecker \cite{ms98} (hereafter MS98) obtained by integrating luminosity dependent infrared spectra of galaxies over their luminosity and redshift distributions. After giving their results on the {$\gamma$-ray}\ optical depth as a function of energy and redshift out to a redshift of 0.3, SD98 applied their calculations by comparing their results with the spectral data on Mrk 421 \cite{mch97} and spectral data on Mrk 501 \cite{ah98}. SD98 make the reasonable simplifying assumption that the IIRF is basically in place at a redshifts $<$ 0.3, having been produced primarily at higher redshifts \cite{mad99}. Therefore SD98 limited their calculations to $z<0.3$. (The calculation of {$\gamma$-ray}\ opacity at higher redshifts \cite{ss97},\cite{ss98} will be discussed in the next section.) SD98 assumed for the IIRF, two of the SEDs given in MS98 \cite{ms98}. Their upper curve now appears to be in better agreement with lower limits from galaxy counts, with Keck telescope, {\it HST}. {\it NICMOS}, {\it ISO} and {\it SCUBA} studies of galaxies at high redshifts (Ref.\cite{mad99} and references therein) and with {\it COBE} data \cite{pug96} - \cite{lag99} (see Figure \ref{irdata}). ~ \begin{figure} \vspace{1.0truecm} \epsfysize=4.0 truein \centerline{\epsfbox{irdata.eps}} \vspace{-5.0truecm} \caption{{The upper infrared SED from Malkan and Stecker compared with observational data and other constraints (courtesy O.C. De Jager).}} \label{irdata} \end{figure} The results of MS98 are also in agreement with upper limits obtained from TeV {$\gamma$-ray}\ studies \cite{sd93} - \cite{bil98}. This agreement is illustrated in Figure \ref{irdata} which shows the upper SED curve from MS98 in comparison with various data and limits. The SD98 results for the absorption coefficient as a function of energy do not differ dramatically from those obtained previously \cite{mp96}, \cite{sd97}; however, they are more reliable because they are based on the empirically derived IIRF given by MS98, whereas all previous calculations of TeV $\gamma$-ray absorption were based on theoretical modeling of the IIRF. The MS98 calculation was based on data from nearly 3000 IRAS galaxies. These data included (1) the luminosity dependent infrared SEDs of galaxies, (2) the 60$\mu$m luminosity function of galaxies and, (3) the redshift distribution of galaxies. \begin{figure} \vspace{2.0truecm} \epsfysize=4.0 truein \centerline{\epsfbox{lowztau.eps}} \vspace{-5.0truecm} \caption{{Optical depth versus energy for {$\gamma$-rays}\ originating at various redshifts obtained using the SEDs corresponding to the lower IIRF (solid lines) and higher IIRF (dashed lines) levels shown in a Figure taken from SD98. As discussed in the text, the higher IIRF curves (dashed lines) are more in line with recent data.}} \label{lowztau} \end{figure} The advantage of using empirical data to construct the SED of the IIRF, as done in MS98, is particularly indicated in the mid-ir range where galaxy observations indicate more flux from warm dust in galaxies than that taken account of in more theoretically oriented models. As a consequence, the mid-IR ``valley'' between the cold dust peak in the far IR and cool star peak in the near IR is partially filled in (see Figure \ref{irdata}). For a source at low redshift, it follows from eq. (1) that {$\gamma$-rays}\ of energy $\sim$ 20 TeV will be absorbed preferentially by photons in the wavelength range of this ``valley'', {\it i.e.}, near 50 $\mu$m. In this range, significant lower limits now exist which are near the predicted IIRF flux (see Figure \ref{irdata}). In fact, the observed flaring spectrum of Mrk 501 has been newly extended to an energy of 24 TeV by observations of the {\it HEGRA} group \cite{kono1}. The new {\it HEGRA} data are well fitted by an $E^{-2}$ source spectrum steepened at energies above a few TeV by intergalactic absorption with the optical depth calculated by SD98 \cite{kono2}. Figure \ref{hegra}, taken from Ref. \cite{kono2}, clearly shows this. The philosophy behind Ref. \cite{kono2} is that the existing lower limits on the mid-ir background flux predict a minimum expected absorption. The derived unabsorbed source spectrum then tells us (1) that there is negligible intrinsic absorption in the source, and (2) the physics of the emission mechanism should give a power-law spectrum with a spectral index of $\sim$2 up to an energy of at least $\sim$ 20 TeV. Consider the source PKS 2155-304, an XBL located at a moderate redshift of 0.117, which has been reported by the Durham group to have a flux above 0.3 TeV of $\sim 4 \times 10^{-11}$ cm$^{-2}$ s$^{-1}$ \cite{chad98}. We predict that this source should have its spectrum steepened by $\sim$ 1 in its spectral index between $\sim 0.3$ and $\sim 3$ TeV and should show an absorption turnover above $\sim 6$ TeV as shown in Figure \ref{2155}. Observations of the spectrum of this source should provide a further test for intergalactic absorption. \section{Absorption of Gamma-Rays at High Redshifts} In order to calculate high-redshift absorption properly, it is necessary to determine the spectral distribution of the intergalactic low energy photon background radiation as a function of redshift as realistically as possible out to frequncies beyond the Lyman limit. This calculation, in turn, requires observationally based information on the evolution of the spectral energy distributions (SEDs) of IR through UV starlight from galaxies, particularly at high redshifts. Salamon and Stecker \cite{ss98} (hereafter SS98) have calculated the {$\gamma$-ray}\ opacity as a function of both energy and redshift for redshifts as high as 3 by taking account of the evolution of both the stellar population spectra and emissivity of galaxies with redshift. In order to accomplish this, they adopted the recent analysis of Fall, {\it et al.}\ \cite{fall96} and also included the effects of metallicity evolution on galactic stellar population spectra. They then gave predicted {$\gamma$-ray}\ spectra for selected blazars and extend our calculations of the extragalactic {$\gamma$-ray}\ background from blazars to an energy of 500 GeV with absorption effects included. \begin{figure} \epsfysize=4.0 truein \centerline{\epsfbox{hegra501.eps}} \vspace{-2.0truecm} \caption{{The bottom curve and points show the new HEGRA data on Mrk 501 in the flaring state \protect\cite{kono1}; the upper line and points show the intrinsic spectrum of the source with the effect of the predicted extragalactic absorption removed \protect\cite{kono2}. }} \label{hegra} \end{figure} \begin{figure} \epsfysize=3.0 truein \hspace{3truecm} \epsfbox{steckerf4rot.eps} \vspace{1truecm} \caption{{Predicted differential absorbed spectrum, for PKS 2155-304 (solid line) assuming an $E^{-2}$ differential source spectrum (dashed line) normalized to the integral flux given in Ref. \protect\cite{chad98} (see text).}} \label{2155} \end{figure} Fall, {\it et al.} \cite{fall96} have devised a method for calculating stellar emissivity which bypasses the uncertainties associated with estimates of poorly defined luminosity distributions of evolving galaxies. The core idea of their approach is to relate the star formation rate directly to the evolution of the neutral gas density in damped {Ly $\alpha$}\ systems, and then to use stellar population synthesis models to estimate the mean co-moving stellar emissivity ${\cal E}_{\nu}(z)$ of the universe as a function of frequency $\nu$ and redshift $z$. The SS98 calculation of stellar emissivity closely follows this elegant analysis, with minor modifications. SS98 also obtained metallicity correction factors for stellar radiation at various wavelengths. Decreased metallicity at high redshifts gives a bluer stellar population spectrum \cite{wo94}, \cite{be94}. The stellar emissivity in the universe is found to peak at $ 1 \le z \le 2$, dropping off steeply at lower redshifts and more slowly at higher redshifts. Indeed, observational data from the Hubble Deep Field to show that metal production has a similar redshift distribution, such production being a direct measure of the star formation rate (see, {\it e.g.}, Ref.\cite{mad99}). With the co-moving energy density $u_{\nu}(z)$ evaluated \cite{ss98} (SS98), the optical depth for {$\gamma$-rays}\ owing to electron-positron pair production interactions with photons of the stellar radiation background can be determined from the expression \cite{sds92} \begin{equation} \label{G} \tau(E_{0},z_{e})=c\int_{0}^{z_{e}}dz\,\frac{dt}{dz}\int_{0}^{2} dx\,\frac{x}{2}\int_{0}^{\infty}d\nu\,(1+z)^{3}\left[\frac{u_{\nu}(z)} {h\nu}\right]\sigma_{\gamma\gamma}(s) \end{equation} where $s=2E_{0}h\nu x(1+z)$, $E_{0}$ is the observed {$\gamma$-ray}\ energy at redshift zero, $\nu$ is the frequency at redshift $z$, $z_{e}$ is the redshift of the {$\gamma$-ray}\ source, $x=(1-\cos\theta)$, and the pair production cross section $\sigma_{\gamma\gamma}$ is zero for center-of-mass energy $\sqrt{s} < 2m_{e}c^{2}$, $m_{e}$ being the electron mass. Above this threshold, \begin{equation} \label{H} \sigma_{\gamma\gamma}(s)=\frac{3}{16}\sigma_{\rm T}(1-\beta^{2}) \left[ 2\beta(\beta^{2}-2)+(3-\beta^{4})\ln\left(\frac{1+\beta}{1-\beta} \right)\right], \end{equation} where $\beta=(1-4m_{e}^{2}c^{4}/s)^{1/2}$. \begin{figure} \vspace{1.0truecm} \epsfysize=3.0 truein \centerline{\epsfbox{opac.eps}} \caption{{ The opacity $\tau$ of the universal soft photon background to {$\gamma$-rays}\ as a function of {$\gamma$-ray}\ energy and source redshift (from SS98) \protect\cite{ss98}. These curves are calculated with and without a metallicity correction. }} \label{opac} \end{figure} Figure \ref{opac} shows the opacity $\tau(E_{0},z)$ for the energy range 10 to 500 GeV, calculated by SD98 both with and without a metallicity correction. Extinction of {$\gamma$-rays}\ is negligible below 10 GeV. The weak redshift dependence of the opacity at the higher redshifts as shown in Figure ~\ref{opac} indicates that the opacity is not very sensitive to the initial epoch of galaxy formation, $z_{max}$. In fact, the uncertainty in the metallicity correction (see Figure \ref{opac}) would obscure any dependence on $z_{max}$ even further. \section{The Effect of Absorption on the Spectra of Blazars and Gamma-ray Bursts} With the {$\gamma$-ray}\ opacity $\tau(E_{0},z)$ calculated out to $z=3$, the cutoffs in blazar {$\gamma$-ray}\ spectra caused by extragalactic pair production interactions with stellar photons can be predicted. Figure \ref{ecritvsz}, based on the results given in Ref. \cite{ss98} (SS98), shows the expected effect of the intergalactic radiation grazar and {$\gamma$-ray}\ burst spectra. This figure plots the critical energy for absorption ({\it i.e.}, for $\tau = 1$) versus redshift. For energies much above the critical energy, the optical depth is greater than 1, leading to a predicted cutoff in the spectrum of the extragalactic source. \begin{figure} \vspace{1.0truecm} \epsfysize=3.5 truein \centerline{\epsfbox{ecritvsz.eps}} \caption{{ The critcal energy for {$\gamma$-ray}\ absorption above which the optical depth is predicted to be greater than 1 as a function of the redshift of the source (from the results of SS98) \protect\cite{ss98} (see text).}} \label{ecritvsz} \end{figure} The discovery of optical and X-ray afterglows of {$\gamma$-ray}\ bursts and the identification of host galaxies with measured redshifts (see, {\it e.g.}, Refs. \cite{me97} and \cite{ku98}) has lead the accumulation of evidence that these bursts are highly relativistic fireballs originating at cosmological distances \cite{liv98} and may be associated primarily with early star formation \cite{dj98}. As indicated in Figure \ref{ecritvsz} {$\gamma$-rays}\ above an energy of $\sim$ 15 GeV will be attenuated if they at emitted at a redshift of $\sim$ 3. On 17 February 1994, the {\it EGRET} telescope observed a {$\gamma$-ray}\ burst which contained a photon of energy $\sim$ 20 GeV \cite{hu94}. As an example, if one adopts the opacity results which include the metallicity correction, the highest energy photon in this burst would be constrained probably to have originated at a redshift less than $\sim$2. Future detectors such as {\it GLAST} Ref. \cite{bl96}, may be able to place better redshift constraints on bursts observed at higher energies. Such constraints may further help to identify the host galaxies of {$\gamma$-ray}\ bursts. Observed cutoffs in grazar spectra may be intrinsic cutoffs in {$\gamma$-ray}\ production in the source, or may be caused by intrinsic {$\gamma$-ray}\ absorption within the source itself. In fact, models of quasar emission can predict natural cutoffs in quasar emission spectra in the relevant energy range above $\sim$ 10 GeV. Whether or not cutoffs in grazar spectra are primarily caused by intergalactic absorption can be determined by observing whether the grazar cutoff energies have the type of redshift dependence predicted here. \section{Acknowledgment} The work presented here was a result of extensive collaboration with O.C. De Jager, M.A. Malkan, and M.H. Salamon, as indicated in the references cited.
\section{Introduction} Let $X$ be a closed oriented surface. Given a Riemannian metric $g$ on $X$, we denote by $\det\Delta_g$ the zeta regularized determinant of the Laplacian $\Delta_g$ associated to $g$. Let $g_0$ be a fixed metric on $X$ and consider the conformal equivalence class $\text{Conf}(g_0)$ of $g_0$. Let $\text{Conf}_0(g_0)$ be the subset of metrics $g\in\text{Conf}(g_0)$ with $\text{Area}(X,g)=\text{Area}(X,g_0)$. In \cite{OPS1}, Osgood, Phillips and Sarnak studied the functional \begin{equation}\label{0.1} h\colon g\in\text{Conf}_0(g_0)\mapsto\det\Delta_g\in\R. \end{equation} One of the main results of \cite{OPS1} states that $h$ has a unique maximum and this maximum is attained at the metric $\tilde g\in\text{Conf}_0(g_0)$ of constant Gauss curvature. As a byproduct, this leads to a new proof of the Riemann uniformization theorem. The present paper grew out of an attempt to generalize this work of Osgood, Phillips and Sarnak to higher dimensions. There exist different possibilities for doing this. In \cite{BCY}, \cite{CY}, for example, Branson, Chang and Yang studied the analogous problem on four-manifolds. More precisley, on a given four-manifold, the authors consider natural, conformally covariant differential operators such as the conformal Laplacian $\Delta+(n-2)s_g/4(n-1)$, and investigate extremals of the zeta function determinants of such operators in a given conformal equivalence class of metrics. We take a different point of view. First of all, the conformal equivalence class $\text{Conf}(g_0)$ determines a unique complex structure on $X$ so that $g_0$ is hermitian with respect to this complex structure. Since $\dim_\C X=1$, $g_0$ is a K\"ahler metric. Let $[\omega_0]\in H^{1,1}(X)$ be the K\"ahler class of $g_0$ and let $\K_{\omega_0}$ be the space of all K\"ahler metrics on $X$ with K\"ahler class equal to $[\omega_0]$. Then $\text{Conf}_0(g_0)=\K_{\omega_0}$. So, we may regard (\ref{0.1}) as a functional on $\K_{\omega_0}$. The variational formula is different from the Polyakov formula, but, of course, it gives the same result as in \cite{OPS1}. This is our starting point for higher dimensional generalizations. We consider a compact K\"ahler manifold $X$ of dimension $n$ and we fix a K\"ahler class $[\omega]\in H^{1,1}(X)$. As above, let $\K_\omega$ be the space of all K\"ahler metrics on $X$ whose K\"ahler class is equal to $[\omega]$. For a given K\"ahler metric $g$, let $T_0(X,g)$ be the Ray-Singer analytic torsion associated to $g$ \cite{RS}. This is a certain weighted product of the regularized determinants of the Dolbeault-Laplace operators $\Delta_{0,q}$, $q=1,...,n$. If $\dim_\C X=1$, then $T_0(X,g)=c(\det\Delta_g)^{1/2}$ for some constant $c\not=0$. Thus, we may regard the functional $$\tau: g\in\K_\omega\mapsto T_0(X,g)\in\R$$ as a higher dimensional analogue of (\ref{0.1}). However since, in general, the harmonic $(0,q)$-forms vary nontrivially for $q>0$, we need to modify this functional appropriately. Let $\parallel\cdot\parallel_{Q,g}$ be the Quillen metric on the determinant line \begin{equation}\label{0.2} \lambda=\bigotimes_{q=0}^n\left(\det H^q(X)\right)^{(-1)^{(q+1)}} \end{equation} \cite{BGS3}. Fix $v\in\lambda$, $v\not=0$, and put $\cQ(X,g):=\parallel v\parallel_{Q,g}$. Then we consider the functional \begin{equation}\label{0.3} \cQ\colon g\in\K_\omega\mapsto \cQ(X,g)\in\R. \end{equation} Now recall that any variation $g_u$, $u\in(-\varepsilon,\varepsilon)$, of $g\in\K_\omega$ is of the form $\omega_u=\omega+\partial\overline\partial\varphi_u$ for some $\varphi_u\in C^\infty(X)$, $u\in(-\varepsilon,\varepsilon)$. Using the results of Bismut, Gillet, and Soul\'e \cite{BGS3}, in section 2 we compute the variation $\delta\cQ/\delta\varphi$ for $\varphi\in C^\infty(X)$. In section 3, we briefly discuss the case of a Riemann surface and the relation with \cite{OPS1}. In section 4 we assume that $X$ admits a metric of constant holomorphic curvature. Such manifolds may be regarded as higher-dimensional analogues of Riemann surfaces. Applying the variational formula of section 2, we show that any metric $g_{KE}\in\K_\omega$ of constant holomorphic curvature is a critical point of $\cQ$. The question, if in this case $g_{KE}$ is the only critical point of $\cQ$, remains open. Also we do not know if $g_{KE}$ is an extremum of $\cQ$. In section 5 we consider $K3$ surfaces. These are examples which do not admit any metric of constant holomorphic sectional curvature. But, on the other hand, by Yau's theorem \cite{Y}, every K\"ahler class $[\omega]$ on a $K3$ surface contains a unique K\"ahler-Einstein metric $g_{KE}$. This metric should be a natural candidate for a critical point of $\cQ:\K_\omega\to\R$. It follows from the variational formula of section 2 that $g_{KE}$ is a critical point of $\cQ$ if and only if the Euler-Lagrange equation \begin{equation}\label{ELE} \Delta c_2(\Omega_{KE})=0 \end{equation} holds. We do not know if there exists any $K3$ surface that admits a K\"ahler-Einstein metric satisfying (\ref{ELE}). However, we can show that there are K\"ahler-Einstein metrics on certain $K3$ surfaces which do not satisfy (\ref{ELE}). In section 6 we introduce a modification of our functional which has a simpler variational formula. For this purpose we assume that $X$ is a complex projective algebraic manifold. Then there exists a positive line bundle $L$ over $X$. We choose the K\"ahler form $\omega$ such that $[\omega]=c_1(L)$. Let $g\in\K_{\omega}$ be the metric corresponding to $\omega$. Following Donaldson \cite{D} we introduce a certain virtual bundle \begin{equation}\label{0.5a} \E=\bigoplus^{4\kappa_2(n+1)}(L-L^{-1})^{\otimes n}\oplus\bigoplus^{-\kappa_1n} (L-L^{-1})^{\otimes (n+1)}, \end{equation} where $\kappa_1$ and $\kappa_2$ are integers which are defined by $\kappa_1=\int_Xc_1(\Omega)\wedge\omega^{n-1}$ and $\kappa_2=\int_X\omega^n$, respectively. Similarly to (\ref{0.2}) there is a determinant line $\lambda(\E)$. Moreover, any hermitian metric $h$ on $L$ determines a hermitian metric $h^\E$ on $\E$. Let $\tilde g\in\K_{\omega}$. With respect to the metrics $(\tilde g,h^\E)$, we form the analytic torsion $T_0(X,\E,\tilde g,h^\E)$ of $X$ with coefficients in $\E$ and the Quillen norm $\parallel\cdot\parallel_{Q,\tilde g,h^\E}$ on $\lambda(\E)$. Fix $v\in\lambda(\E)$, $v\not=0$, and set $$\cQ_L(\tilde g,h):=\parallel v\parallel_{Q,\tilde g,h^\E}.$$ This functional of $(\tilde g,h)$ can be turned into a functional on $\K_{\omega}$ as follows. According to our choice of $\omega$, there exists a hermitian metric $h$ on $L$ such that the curvature $\Theta_{h}$ of $h$ satisfies $\Theta_{h}=-2\pi i\omega$. Then $h$ is determined by $g$ up to multiplication by a positive scalar. Let $$\H_{\omega}=\{\varphi\in C^\infty(X)\mid \omega+i\partial\overline\partial\varphi>0\}$$ be the space of K\"ahler potentials. Given $\varphi\in\H_\omega$, let $\omega(\varphi)=\omega+i\partial\overline\partial\varphi$ and let $g(\varphi)$ be the metric with K\"ahler form $\omega(\varphi)$. It is well known that the map $\varphi\in\H_\omega\mapsto g(\varphi)\in\K_\omega$ is surjective. Furthermore, let $h(\varphi)=e^{-2\pi\varphi}h$. Then we put $$\cQ_L(\varphi)=\cQ_L(g(\varphi),h(\varphi))^{(-1)^n}, \quad\varphi\in\H_\omega.$$ Using the variational formula established in \cite{BGS3}, one can show that the functional $\cQ_L$ satisfies $\cQ(\varphi+c)=\cQ(\varphi)$ for all $c\in\R$ and hence, it can be pushed down to a functional on $\K_\omega$. By the same variational formula one can compute the variation of $\cQ_L$. Let $\varphi_t$, $t\in(a,b)$, be a smooth path in $\H_\omega$ and set $g_t=g(\varphi_t)$. Then \begin{equation}\label{0.4} \frac{\partial\cQ_L}{\partial t}(g_t)=c_n\int_X\dot\varphi_t(s(\varphi_t)-s_0)\omega(\varphi_t)^n, \end{equation} where $s(\varphi_t)$ is the scalar curvature of the metric $g(\varphi_t)$, $s_0$ is the normalized total scalar curvature and $c_n>0$ is a certain constant that depends only on $n$ and $[\omega]$. It follows from (\ref{0.4}) that the critical points of $\cQ_L$ are exactly the metrics $g\in\K_\omega$ of constant scalar curvature. Any such metric is an extremal metric in the sense of Calabi \cite{Ca}. Furthermore, formula (\ref{0.4}) agrees, up to the constant $c_n$ and the sign, with the variation of the K-energy $\mu\colon \K_\omega\to\R$ introduced by Mabuchi \cite{Ma1}. The K-energy has been studied by Bando and Mabuchi \cite{B}, \cite{BM}, \cite{Ma2}, mainly in connection with Futaki's obstruction to the existence of K\"ahler-Einstein metrics in the case $c_1(X)>0$. In \cite{Ma2}, Mabuchi defined a natural Riemannian structure on $\K_\omega$. He proved that the sectional curvature of $\K_\omega$ is nonpositive and that Hess($\mu$) is positive semidefinite everywhere. Therefore $\cQ_L$ is also a convex functional. But this is not sufficient to determine the type of the critical point. In section 7 we assume that $\K_\omega$ contains a K\"ahler-Einstein metric. Then we can say more about the critical points of $\cQ_L$. The main result is the following theorem. \begin{thm}\label{th0.1} Let $(X,L)$ be a polarized projective algebraic manifold. Choose a K\"ahler form $\omega$ on $X$ such that $[\omega]=c_1(L)$. Assume that $\K_\omega$ contains a K\"ahler-Einstein metric $g_{KE}$. If $c_1(X)\le0$, then $\cQ_L$ has a unique maximum which is attained at $g_{KE}$. If $c_1(X)>0$, then $\cQ_L$ attains its absolute maximum on the subset $\K_{KE}\subset\K_\omega$ of K\"ahler-Einstein metrics. \end{thm} To prove the first part of Theorem \ref{th0.1}, we use the evolution of the metric by the complex analogue of Hamilton's Ricci flow equation \begin{equation}\label{0.5} \frac{\partial \tilde g_{i\overline j}}{\partial t}=-\tilde r_{i\overline j}+\frac{s_0}{2n}\tilde g_{i\overline j},\quad \tilde g_{i\overline j}(0)=g_{i\overline j}, \end{equation} where $\tilde r _{ij}$ is the Ricci tensor of $\tilde g$ and $s_0$ the normalized total scalar curvature. Cao \cite{C} proved that (\ref{0.5}) has a unique solution $\tilde g_{i\overline j}(t)$ which exists for all time and, furthermore, if $c_1(X)\le0$, then $\tilde g(t)$ converges to $g_{KE}$ as $t\to\infty$. Now the main observation is that the K-energy decreases along the flow generated by (\ref{0.5}). The case $c_1(X)>0$ is more complicated. First of all, there are obstructions to the existence of K\"ahler-Einstein metrics \cite{Fu}, \cite{Ti2}. Moreover, the subspace $\K_E\subset \K_\omega$ of K\"ahler-Einstein metrics may have positive dimension. If $\K_E\not=\emptyset$, it was proved in \cite{BM} and \cite{B} that $\mu$ takes its minimum on $\K_E$. This implies the second statement of the theorem. Theorem \ref{th0.1} has some implication for the spectral determination of K\"ahler-Einstein metrics. As above, let $g$ be a K\"ahler metric on $X$ such that $[\omega_g]=c_1(L)$ and pick a hermitian metric $h$ on $L$ such that $\Theta_h=-2\pi i\omega_g$. Recall that $h$ is uniquely determined by $g$ up to multiplication by a positive constant. Let $q\in\{0,...,n\}$ and $k\in\N$. The Dolbeault-Laplace operator in $\Lambda^{0,q}(X,L^{\otimes k})$, associated to $(g,h)$, remains unchanged if we multiply $h$ by $\lambda\in\R^+$. Therefore, it is uniquely determined by $g$ and will be denoted by $\Delta^g_{0,q,k}$. Let Spec($\Delta_{0,q,k}$) denote the spectrum of $\Delta^g_{0,q,k}$. By the above, the spaces $\H^{0,q}(X,L^{\otimes k})$ of $L^{\otimes k}$-valued harmonic $(0,q)$-forms also depend only on $g$. Although the inner product in $\H^{0,q}(X,L^{\otimes k})$ associated to $(g,h)$ depends on $h$, the induced $L^2$-norm on $\lambda(\E)$ is invariant under multiplication of $h$ by positive scalars and therefore, depends only on $g$. We denote it by $\parallel\cdot \parallel_g$. \begin{cor}\label{cor0.2} Let $(X,L)$ be a polarized projective algebraic manifold and suppose that $c_1(X)\le0$. Choose a K\"ahler form $\omega$ on $X$ such that $[\omega]=c_1(L)$. Let $g_{KE}$ be the unique K\"ahler-Einstein metric in $\K_\omega$. Let $g\in\K_\omega$ and suppose that the following holds: \begin{enumerate} \item[1)]$\Spec(\Delta^g_{0,q,k}) =\Spec(\Delta^{g_{KE}}_{0,q,k})$ for all $q=0,...,n$ and $k=-(n+1),...,n+1$. \item[2)]$\parallel\cdot\parallel_g=\parallel\cdot\parallel_{g_{KE}}$. \end{enumerate} Then $g=g_{KE}$. \end{cor} It seems to be likely that the corollary can be improved. One would expect that $g_{KE}$ is already uniquely determined by the spectra of the Dolbeault-Laplace operators. In the final section 8 we briefly discuss the relation to moduli spaces. These are slight modifications of results due to Fujiki and Schumacher \cite{FS}. First we consider a metrically polarized family of compact Hodge manifolds $(\pi:\X\to S,\tilde\omega,\L)$. Using $\L$, we introduce the family version of the virtual bundle (\ref{0.5a}) which we also denote by $\E$. Associated to $\E$ is the determinant line bundle $\lambda(\E)$ on $S$ equipped with the Quillen metric $h^{\lambda(\E)}$. If the family $(\pi:\X\to S,\tilde\omega,\L)$ is effective, then by Fujiki and Schumacher \cite{FS}, one can define the generalized Weil-Petersson metric $\widehat{h}_{WP}$ on $S$ and it follows that the first Chern form $c_1(\lambda(\E),h^{\lambda(\E)})$ of the determinant line bundle satisfies $$c_1(\lambda(\E),h^{\lambda(\E)})=a_n\widehat\omega_{WP}$$ where $a_n=-2^{n+1}\kappa_2(n+1)!$. This result extends to the moduli space ${\mathfrak M}_{H,e}$ of extremal Hodge manifolds \cite{FS}. \smallskip \noindent {\bf Acknowledgement.} The authors would like to thank Kai K\"ohler for some very helpful comments and remarks. \section[Preliminaries]{Preliminaries} \setcounter{equation}{0} Let $(X,g)$ be a compact K\"ahler manifold of complex dimension $n$. We denote by $\omega=\omega_g$ the K\"ahler form of $g$. Then the volume form of $g$ is given by $dv_g=\omega^n_g/n!$. We always choose the holomorphic connection on $X$. The Riemann and Ricci curvature tensors will be denoted by $R$ and $r$, respectively, and the scalar curvature by $s_g$. In terms of the Ricci tensor $r$, the Ricci form $\rho$ is defined by $$\rho(\varphi,\psi)=r(J\varphi,\psi),$$ where $J$ denotes the complex structure of $X$. For a K\"ahler manifold, $i\rho$ is the curvature of the canonical line bundle $K=(\Lambda^{n}T^{1,0}X)^*$. This has important implications for the scalar curvature $s_g=\operatorname{Tr}(r)$. It can be calculated by the formula \begin{equation}\label{1.1} s\thinspace\omega^n=2n\;\rho\wedge\omega^{n-1}. \end{equation} Since $\rho/2\pi$ represents the first Chern class $c_1(X)=c_1(T^{1,0}X) =-c_1(K)$, it follows that the average value \begin{equation}\label{1.2} s_0=\frac{\int_X s_g dv_g}{\int_X dv_g}=4n\pi\frac{\int_X c_1(\Omega)\wedge\omega^{n-1}}{\int_X\omega^n} \end{equation} of the scalar curvature is a topological invariant, depending only on the K\"ahler class $[\omega]$. \smallskip \noindent Let $\E\to X$ be a holomorphic vector bundle with hermitian metric $h^\E$. Let \begin{equation*} 0\to \Lambda^{0,0}(X,\E)\stackrel{\overline\partial}{\longrightarrow}\Lambda^{0,1}(X,\E)\stackrel{\overline\partial}{\longrightarrow} \cdots\stackrel{\overline\partial}{\longrightarrow}\Lambda^{0,n}(X,\E)\to 0 \end{equation*} be the Dolbeault complex and let $\Delta^\E_{0,q}= \overline\partial\thinspace \overline\partial^* +\overline\partial^*\overline\partial$ be the corresponding Dolbeault-Laplace operator acting in $\Lambda^{0,q}(X,\E)$. Let $P^\E_{0,q}$ be the projection on harmonic $(0,q)$-forms and for $\Re(s)>n/2$, let \begin{equation*} \zeta_q(s;\E)=\frac{1}{\Gamma(s)}\int^\infty_0t^{s-1}\mbox{Tr} (e^{-t\Delta^\E_{0,q}}-P^\E_{0,q})dt \end{equation*} be the zeta function of $\Delta^\E_{0,q}$. It admits a meromorphic continuation to $\C$, also denoted by $\zeta_q(s;\E)$, which is regular at $s=0$. Then the zeta regularized determinant $\det\Delta^\E_{0,q}$ of $\Delta^\E_{0,q}$ is defined by \begin{equation*} \det\Delta^\E_{0,q}=\exp\left(-\frac{d}{ds}\zeta_q(s;\E)\big|_{s=0}\right). \end{equation*} \smallskip \noindent The Ray-Singer analytic torsion is defined to be \begin{equation*} T_0(X,\E)=\left(\prod^n_{q=0} (\det\Delta^\E_{0,q})^{(-1)^{q+1} q}\right)^{1/2} \end{equation*} \cite{RS}. This number depends, of course, on the metrics $(g,h^\E)$ and if it is necessary to indicate this dependence we shall write $T_0(X,\E,g,h^\E)$ for the analytic torsion. Let \begin{equation*} \lambda(\E)=\bigotimes^n_{q=0} (\det H^q(X,\E))^{(-1)^{q+1}} \end{equation*} be the determinant line associated to the Dolbeault complex. Let $\H^{0,q}(\E)$ be the space of $\E$-valued harmonic $(0,q)$-forms. For each $q$, there is a canonical isomorphism ${\H}^{0,q}({\E})\stackrel{\cong}{\longrightarrow}H^q(X,\E)$. Thus, using the $L^2$-metric on ${\H}^{0,q}(\E)$, we get a metric $\parallel\cdot\parallel_{L^{2}}$ on $\lambda(\E)$. The Quillen metric on $\lambda(\E)$ is then defined by \begin{equation}\label{1.3} \parallel\cdot\parallel_Q=\parallel\cdot\parallel_{L^{2}} \cdot T_0(X,\E). \end{equation} We shall also consider virtual holomorphic bundles $\E=\sum_{k=1}^mn_k\E_k$, where $n_k\in\Z$ and $\E_k\to X$ are holomorphic vector bundles. For such $\E$ we set \begin{equation*} \lambda(\E)=\bigotimes_{k=1}^m\lambda(\E_k)^{n_k}. \end{equation*} Here for $n_k<0$, $\lambda(\E_k)^{n_k}:=(\lambda(\E_k)^*)^{-n_k}.$ If each $\E_k$ is equipped with a hermitian metric, we get a metric $\parallel\cdot\parallel_Q$ on $\lambda(\E)$ which is the tensor product of the induced Quillen metrics on $\lambda(\E_k)^{n_k}$. \smallskip \noindent Let $v\in \lambda(\E)$, $v\neq 0$. Then we define \begin{equation*} {\mathcal Q}(X,\E,g,h):=\parallel v\parallel_{Q,g,h}. \end{equation*} If $X$ and $\E$ are fixed, we shall write $\cQ(g,h)$ in place of $\cQ(X,\E,g,h)$. Let $\omega$ be a fixed K\"ahler form. Put $$\K_{\omega}=\bigl\{g\mid g\text{ K\"ahler metric on } X\text{ with }[\omega_g]=[\omega]\bigr\}. $$ Let \begin{equation}\label{1.4} \H_{\omega}=\bigl\{\varphi\in C^\infty(X,\R)\mid \omega+i\partial\overline \partial\varphi>0\bigr\} \end{equation} be the corresponding space of K\"ahler potentials. If $\varphi\in\H_{\omega}$, then \begin{equation}\label{1.5} \omega(\varphi):=\omega+i\partial\overline\partial\varphi \end{equation} is the K\"ahler form of a K\"ahler metric $g(\varphi)\in\K_\omega$ and it is well known that the natural map \begin{equation}\label{1.6} \varphi\in\H_\omega\to g(\varphi)\in\K_\omega \end{equation} is surjective. If $\varphi\in\H_\omega$ then for each $c\in\R$, one has $\varphi+c\in\H_\omega$ and $\omega(\varphi)=\omega(\varphi+c)$. On the other hand, if $\varphi,\psi\in\H_\omega$ are such that $\omega(\varphi)=\omega(\psi)$ then $\partial\overline\partial(\varphi-\psi)=0$ and therefore $\varphi-\psi$ is constant. Let $$\H_{\omega}^0=\left\{\varphi\in\H_\omega\;\Big|\; \int_X\varphi\omega^n=0\right\}.$$ Then the restriction of the map (\ref{1.6}) to the subspace $\H^0_\omega$ induces an isomorphism \begin{equation} \H_\omega^0\cong\K_\omega. \end{equation} Furthermore, for any $\varphi\in C^\infty(X)$ there exists $\epsilon>0$ such that \begin{equation*} \omega_u=\omega+iu\partial\overline\partial\varphi>0 \end{equation*} for all $u\in\R,|u|<\epsilon$. The corresponding family $g_u,|u|<\epsilon$, of K\"ahler metrics will be called the variation of $g$ in the $\varphi$-direction. \section[The variational formala]{The variational formula} \setcounter{equation}{0} Given a smooth family $g_t$, $t\in \R$, of K\"ahler metrics, we will denote by $*_t$ the Hodge star operator with respect to $g_t$ and we set \begin{equation}\label{2.1} U_t:= (g_t)^{-1} \frac{d}{dt} (g_t), \quad \alpha_t:=*^{-1}_t \frac{d}{dt} (*_t). \end{equation} \noindent We observe that if $g_t$ is the variation of $g$ in the $\varphi$-direction, then we have \begin{equation}\label{2.2} \mbox{Tr}(U_t)= \sum_{\alpha,{\overline\beta}} g^{\alpha{\overline\beta}}_t \frac{\partial^2\varphi}{\partial z_\alpha \partial{\overline z}_\beta} =-\frac{1}{2}\Delta^{t}_{0,0}\varphi. \end{equation} The variation of ${\mathcal Q}(X,\E,g,h)$ has been computed by Bismut, Gillet and Soul\' e \cite[Theorem 1.22]{BGS3}. We recall their result. Let $g_t,t\in\R$, be a smooth family of K\"ahler metrics on $X$ and $h_t,\; t\in\R$, a smooth family of hermitian metrics on $\E$. Let $\Omega_t$ and $L^\E_t$ be the curvature of the Hermitian holomorphic connection on $(T^{(0,1)} X,g_t)$ and $(\E,h_t)$, respectively. Let Td be the Todd genus and $Td_j$ its $j$-th component. The Todd genus is normalized as in \cite{BGS3}. Then \begin{equation}\label{2.3} \begin{split} \frac{\partial}{\partial t} \log {\mathcal Q}(X,\E,g_t,h_t) & = \frac{1}{2}\left( \frac{1}{2\pi i}\right)^n\int_X\frac{\partial}{\partial b} \bigg[ \mbox{Td}(-\Omega_t-b U_t) \\ & \times \mbox{Tr} \left( \exp(-L^\E_t -b(h^\E_t)^{-1} \frac{\partial h^\E_t}{\partial t})\right)\bigg]_{b=0}. \end{split} \end{equation} In particular, if we assume that $\E$ is trivial, this formula is reduced to \begin{equation}\label{2.4} \frac{\partial}{\partial t}\log {\mathcal Q}(X,g_t)= \frac{1}{2}\left(\frac{1}{2\pi i}\right)^n \int_X \frac{\partial}{\partial b} \left[ \mbox{Td}_{n+1}(-\Omega_t-bU_t)\right]\Big|_{b=0}. \end{equation} Using the definition of ${\mathcal Q}(X,g)$ and (1.120) of \cite{BGS3}, it follows that \begin{equation}\label{2.5} \frac{\partial}{\partial t}\log{\mathcal Q}(X,g_t)= \frac{\partial}{\partial t}\log T_0(X,g_t)-\frac{1}{2} \sum^n_{q=0}(-1)^q\operatorname{Tr}(\alpha_t P^t_{0,q}). \end{equation} For the application that we have in mind we need a more explicit version of the variational formula. Recall that the Todd genus can be expressed in terms of the Chern classes. Therefore, we have to compute the corresponding derivatives of the Chern classes which are also normalized as in \cite{BGS3}. \begin{lem}\label{l2.1} For all $j\ge1$, we have \begin{equation}\label{2.6} \frac{\partial}{\partial b}\left( {c_j}(-\Omega-b U)\right)\Big|_{b=0}= \sum^{j-1}_{k=0}(-1)^{j+k}\operatorname{Tr}(\Omega^k U) {c}_{j-k-1}(\Omega). \end{equation} \end{lem} \begin{pf} We proceed by induction on $j$. For $j=1$ we have \begin{equation*} \frac{\partial}{\partial b}\left( c_1(-\Omega-b U)\right)\Big|_{b=0} = -\mbox{Tr}(U) \end{equation*} proving (\ref{2.6}) in this case. Now suppose that (\ref{2.6}) holds for $j$. We express the Chern classes by the power sums $s_k$. If we formally write \begin{equation*} \sum^n_{j=0} c_j(\Omega)x^j=\prod^n_{j=1}(1+\gamma_j x), \end{equation*} then $s_k$ is the $k$-th elementary symmetric function in $\gamma_1,\ldots,\gamma_n$. Using Newton's formula, we obtain \begin{equation*} c_{j+1}(-\Omega-b U)=\frac{1}{j+1}\sum^{j+1}_{k=1} (-1)^{k+1} (c_{j+1-k} s_k)(-\Omega-b U). \end{equation*} Since $s_k(R)=\mbox{Tr}(R^k)$, we get \begin{equation*} \frac{\partial}{\partial b}(s_k(-\Omega -b U))\mid_{b=0} =k(-1)^k\mbox{Tr}(\Omega^{k-1}U). \end{equation*} Using the induction hypothesis, we deduce \begin{equation*} \begin{split} \frac{\partial}{\partial b}\bigl( &{c}_{j+1}(-\Omega-b U)\bigr) \Big|_{b=0}\\ &=\frac{1}{j+1}\sum^{j+1}_{k=1} \left\{ \sum^{j-k}_{l=0}(-1)^{j+l} \mbox{Tr}(\Omega^lU) c_{j-k-l}(\Omega) s_k(-\Omega)\right.\\ &\hskip130pt -k{ c}_{j+1-k}(-\Omega)\mbox{Tr}(\Omega^{k-1} U)\Biggr\}\\ & =\frac{1}{j+1} \sum^j_{l=0} \mbox{Tr}(\Omega^l U) \left\{ \sum^{j-l}_{k=1} (-1)^{l+j+k} { c}_{j-k-l}(\Omega)s_k(\Omega)\right.\\ &\hskip143pt -(-1)^{j+l}(l+1){ c}_{j-l}(\Omega)\Biggr\}\\ &= \sum^j_{l=0} (-1)^{j+1+l}\mbox{Tr}(\Omega^l U) {c}_{j-l}(\Omega). \end{split} \end{equation*} \end{pf} Expressing the Todd genus in terms of the Chern classes and using Lemma \ref{l2.1}, it follows from (\ref{2.4}) that there exist $a_{j_{0}\cdots j_{n+1}}\in\C$, $j_0,\ldots,j_{n+1}\in\N$, such that for all $\varphi\in C^\infty(X):$ \begin{equation}\label{2.7} \begin{split} \frac{\partial}{\partial t}\log{\mathcal Q}(X,g_t) &=\sum_{j_0+\cdots+j_{n+1}=n} a_{j_{0}\cdots j_{n+1}}\\ &\quad\times\int_X \operatorname{Tr}(\Omega^{j_{n+1}}_t U_t) {c}_{j_{0}}(\Omega_t)\cdots {c}_{j_{n}}(\Omega_t). \end{split} \end{equation} The coefficients $a_{ j_{0}\cdots j_{n+1}}\in\C$, $j_0,\ldots,j_{n+1}\in N$, can be computed explicitly. For example, if $n=2$, then formula (\ref{2.7}) takes the form \begin{equation}\label{2.8} \begin{split} \frac{\partial}{\partial t}\log{\mathcal Q}(X,g_t)& =\frac{1}{48}\int_X\operatorname{Tr}(\Omega_tU_t)c_1(\Omega_t)\\ &-\frac{1}{48}\int_X\operatorname{Tr}(U_t)(c_2(\Omega_t)+c_1(\Omega_t)^2). \end{split} \end{equation} Let $\varphi\in C^\infty(X)$ and let $g_u$, $|u|<\epsilon$, be the variation of $g$ in the direction of $\varphi$. Put \begin{equation}\label{2.9} \frac{\delta}{\delta\varphi}\log{\mathcal Q}(X,g)=\frac{\partial}{\partial u} \log {\mathcal Q}(X,g_u)\Big|_{u=0}. \end{equation} If $n=2$, it folows from (\ref{2.2}) and (\ref{2.8}) that \begin{equation}\label{2.10} \begin{split} \frac{\delta}{\delta\varphi}\log{\mathcal Q}(X,g)& =\frac{1}{48}\int_X\operatorname{Tr}(\Omega U_0)c_1(\Omega)\\ &+\frac{1}{96}\int_X\Delta\varphi(c_2(\Omega)+c_1(\Omega)^2). \end{split} \end{equation} \section[Riemann surfaces]{Riemann surfaces} \setcounter{equation}{0} In this section we consider the case $n=1$. Then $X$ is a compact oriented surface without boundary. First observe that \begin{equation*} \alpha(1)=-\frac{\Delta\varphi}{2}\; \mbox{and}\; \alpha({d\overline z})=0. \end{equation*} Hence, it follows from (\ref{2.5}) that \begin{equation*} \frac{\delta}{\delta\varphi}\log{\mathcal Q}(X,g)=\frac{\delta}{\delta\varphi}\log T_0(X,g). \end{equation*} Furthermore, $\overline\partial: \wedge^{0,0}(X)\to\wedge^{0,1}(X)$ is an isomorphism on nonzero eigenspaces. This implies that $\det \Delta_{0,1}=\det\Delta_{0,0}$. Let $\Delta_g=d^* d$ denote the Laplacian on functions. Then $\Delta_g=2\Delta_{0,0}$ and therefore we get \begin{equation}\label{3.1} 2\log T_0(X,g)=\log \det\Delta_g+(\log 2)\left( 1-\frac{\chi(X)}{6}\right). \end{equation} Thus there exists $c>0$ such that \begin{equation*} T_0(X,g)=c(\det\Delta_g)^{1/2}. \end{equation*} Next we describe the space ${\mathcal K}_\omega$ for a given K\"ahler class $[\omega ]\in H^{1,1}(X)$. Since $H^{1,1}(X)\cong\R,\; [\omega]$ is unique up to multiplication by $\R^+$. Let $g$ be the metric with K\"ahler form $\omega$ and let $\Delta=\Delta_g$. Let $\psi\in C^\infty(X)$. Then $\omega+i\partial\overline\partial \psi>0$ holds if and only if $1-\Delta\psi >0$. Thus the space $\H_\omega$, which is defined by (\ref{1.4}), is given by \begin{equation}\label{3.2} {\mathcal H}_\omega \cong \{ \psi\in C^\infty(X)\mid 1-\Delta\psi>0\}. \end{equation} Then by using $g(\psi)=(1-\Delta\psi)g$, we may regard $T_0(X,g)$ as a functional on ${\mathcal H}_\omega$ and by (\ref{3.1}) we have \begin{equation*} 2 \log T_0(X,\psi)=\log \det\Delta_{(1-\Delta\psi)g}+C,\quad \psi\in{\mathcal H}_\omega. \end{equation*} Let \begin{equation*} \mbox{Conf}(g)=\{e^{2\varphi}g\mid\varphi\in C^\infty(X)\} \end{equation*} be the conformal equivalence class of metrics on $X$ which contains $g$. Let \begin{equation*} \mbox{Conf}_0(g)=\{{\widetilde g}\in\mbox{Conf}(g)\mid \mbox{Area}(X,{\widetilde g})=\mbox{Area}(X,g)\}. \end{equation*} \begin{lem}\label{l3.1} Let $g$ be a K\"ahler metric on $X$ with K\"ahler form $\omega$. Then \begin{equation*} {\mathcal K}_\omega=\Cf_0(g). \end{equation*} \end{lem} \begin{pf} Let ${\widetilde g}\in{\mathcal K}_\omega$. By (\ref{3.2}) there exists a unique $\psi\in C^\infty(X)$ with $\int\psi\;dv_g=0$ such that ${\widetilde g}=(1-\Delta\psi)g$. Since $1-\Delta\psi>0$, there exists $\varphi\in C^\infty(X)$ such that $e^{2\varphi}=1-\Delta\psi$. Moreover \begin{equation*} \int_X e^{2\varphi}dv_g=\int_Xdv_g-\int_X \Delta\psi dv_g=\int_Xdv_g. \end{equation*} Hence there exists a unique $\varphi\in C^\infty(X)$ such that ${\widetilde g}=e^{2\varphi}g\in{\mbox{Conf}}_0(g)$. On the other hand, let ${\widetilde g}\in\mbox{Conf}_0(g)$. Then ${\widetilde g}=e^{2\varphi}g$ and $\int_X(e^{2\varphi}-1)dv_g=0$. Therefore, there exists a unique $\psi\in C^\infty(X)$ with $\int\psi\;dv_g=0$ such that $e^{2\varphi}=1-\Delta\psi>0$. Hence \begin{equation*} {\widetilde g}=(1-\Delta\psi)g\in{\mathcal K}_\omega. \end{equation*} \end{pf} \noindent The functional \begin{equation}\label{3.3} {\widetilde g}\in\mbox{Conf}_0(g)\longmapsto\log\det\Delta_{ {\widetilde g}}\in\R \end{equation} has been studied by Osgood, Philipps and Sarnak \cite{OPS1}. By (\ref{3.1}) and Lemma 3.1, we see that, up to a constant, (\ref{3.3}) coincides with ${\widetilde g}\in{\mathcal K}_\omega\longmapsto \log{\mathcal Q}(X,{\widetilde g})$. We shall now derive the main result of \cite{OPS1} by our approach. \smallskip \noindent Let $g_t\in\K_\omega$, $t\in\R$, be a smooth family of metrics. Then there exists a smooth family $\psi_t\in C^\infty(X)$ such that $g_t=g(\psi_t)$. Using Lemma 2.1, the variational formula (\ref{2.4}) gives \begin{equation}\label{3.4} \frac{\partial}{\partial t}\log{\mathcal Q}(X,g_t)=-\frac{1}{24}\int_X c_1(\Omega_t)\Delta_t\dot \psi_t=\frac{1}{12\pi}\int_X K_t\Delta_t\dot \psi_t \;dv_t, \end{equation} where $K_t$ is the Gauss curvature of $g_t$. We note that the last equality follows from $$c_1(\Omega_t)=-\frac{2K_t}{\pi}dv_t.$$ In particular, we get \begin{equation}\label{3.4a} \frac{\delta}{\delta\psi}\log{\mathcal Q}(X,g)=-\frac{1}{24}\int_X c_1(\Omega)\Delta \psi=\frac{1}{12\pi}\int_X K\Delta \psi\; dv. \end{equation} \smallskip \begin{remark} This formula should be compared with the corresponding variational formula (1.12) in \cite{OPS1}. Given $\psi\in C^\infty(X)$, there exists $\epsilon>0$ such that $1-u\Delta \psi>0$ for $|u|< \epsilon$. Then by Lemma \ref{l3.1}, there exists $\phi_u\in C^\infty(X)$ such that \begin{equation}\label{3.5} 1-u\Delta\psi=e^{2\phi_u},\quad |u|<\epsilon. \end{equation} The family $g_u=e^{2\phi_u}g,\quad |u|<\epsilon$, of metrics is a deformation of $g$ in $\mbox{Conf}_0(g)$. By (\ref{3.5}), we have \begin{equation*} \phi_0=0,\quad {\dot\phi}_0=-\frac{\Delta\psi}{2}. \end{equation*} Then with respect to the variation defined by $g_u$, in formula (1.12) of \cite{OPS1} we have $\delta\varphi={\dot\phi}_0$, $\varphi=0$ and $\delta \log A=0$. Hence, up to a constant, (\ref{3.4a}) corresponds to (1.12) of \cite{OPS1}. \end{remark} By (\ref{3.4a}), a metric ${\widetilde g}$ is critical for $\log{\mathcal Q}(X,g)$ if and only if its Gauss curvature $K$ is constant. To investigate the critical point, we use the evolution of a Riemannian metric $g$ on $X$ by Hamilton's Ricci flow equation with $s_0$ as in (\ref{1.2}) \begin{equation}\label{3.6} \frac{\partial g_{ij}}{\partial t}=(s_0-s_g)g_{ij}. \end{equation} Note that by the Gauss-Bonnet formula $$s_0=\frac{4\pi\chi(X)}{A},$$ where $A=\text{Area}(X,g)$. In \cite{Ha}, Hamilton proved that if $s_0\le0$ or if $s_g>0$, then for any initial data, (\ref{3.6}) has a unique solution $g(t)$ which exists for all time and as $t\to\infty$, converges to a metric of constant curvature. The case $s_0>0$ was completed by Chow \cite{Ch} who proved that for $s_0>0$, the scalar curvature $s(t)$ of the solution $g(t)$ of (\ref{3.6}) becomes positive in finite time. Now if $g(0)$ is hermitian then $g(t)$ is hermitian for all time. Furthermore, the Ricci flow is area preserving \cite[p.238]{Ha}. Therefore, the Ricci flow preserves the space $\K_\omega$. Hence, there exists $u(t)\in C^\infty(X)$ which is a smooth function of $t\in\R^+$ such that, with respect to a local holomorphic parameter $z$, $$g(t)=g(0)+\frac{\partial^2u(t)}{\partial z\partial\overline{z}}dz\otimes d\overline{z}.$$ Let $g(t)=h(t)dz\otimes d\overline{z}$. Then (\ref{3.6}) implies \begin{equation}\label{3.7} \Delta_t \dot u=-h^{-1}\frac{\partial^2\dot u}{\partial z\partial\overline{z}} =(s_g-s_0). \end{equation} Combined with (\ref{3.4}) we obtain \begin{equation}\label{3.8} \begin{split} \frac{\partial}{\partial t}\log\cQ(g(t))&=\frac{1}{24\pi}\int_Xs(t)\Delta_t\dot u(t)\;dv_t\\ &=\frac{1}{24\pi}\int_X(s(t)-s_0)\Delta_t\dot u(t)\;dv_t\\ &=\frac{1}{24\pi}\int_X(s(t)-s_0)^2\;dv_t\ge0, \end{split} \end{equation} where $s(t)$ denotes the scalar curvature of the metric $g(t)$. If $s_0\le0$, Hamilton \cite{Ha} proved that there exist $\varepsilon>0$ and $C>0$ such that \begin{equation}\label{3.10} |s(t)-s_0|\le Ce^{-\varepsilon t},\quad t\in\R^+. \end{equation} If $s_0>0$, he established the same estimate under the additional assumption that $s_g>0$. By Chow's results \cite{Ch} it follows that (\ref{3.10}) holds in the case $s_0>0$ too. Hence we can integrate (\ref{3.8}) and we get $$\log\cQ(g(\infty))-\log\cQ(g(0))=\frac{1}{24\pi} \int_0^\infty\int_X(s(t)-s_0)^2\;dv_t\;dt\ge0.$$ Furthermore, if $s(0)$ is not constant then $(s(t)-s_0)^2>0$ for $t$ in a neighborhood of $0$ and therefore, the above inequality is strict. Thus we proved the following theorem. \begin{thm} $\cQ$ has a unique maximum on $\K_\omega$ which is attained at the metric $g^*\in\K_\omega$ of constant curvature. \end{thm} This result was first proved by Osgood, Phillips and Sarnak in \cite{OPS1} by a different method. \section[Constant holomorphic curvature]{Manifolds with constant holomorphic curvature} \setcounter{equation}{0} Let $(X,g)$ be a K\"ahler manifold and let $J$ denote the complex structure of $X$. Recall that $(X,g)$ is said to have constant holomorphic curvature $H\in\R$, if for any $z\in X$ and any $\xi\in T_z X$ with $\langle \xi,\xi\rangle=1$, the holomorphic curvature \begin{equation*} H_z(\xi)=\langle R_z(\xi,J\xi)\xi,J\xi\rangle \end{equation*} is equal to $H$. We note that any metric of constant holomorphic curvature is K\"ahler-Einstein \cite[(6.12)]{Go}. Furthermore, compact manifolds with constant holomorphic curvature $H$ can be classified as follows \cite{Be}: If $H>0$, then $X$ is isomorphic to $\C P^n$ equipped with the rescaled Fubini-Study metric; if $H=0$, then $X$ is isomorphic to a complex torus $\Gamma\backslash \C^n$ with the flat metric; if $H<0$ then $X$ is isomorphic to a compact quotient $\Gamma\backslash D^n$ of the complex unit ball $D^n\subset \C^n$, equipped with a Bergmann metric, by a discrete group $\Gamma$ of isometries. Thus manifolds with constant holomorphic curvature may be regarded as higher-dimensional analogues of Riemann surfaces. So, if we think of generalizing the results of Osgood, Philipps and Sarnak \cite{OPS1} to higher dimensions, these are the most obvious candidates for being critical points of the functional $\log{\mathcal Q}(X,g)$. \smallskip \noindent We wish to apply the variational formula (\ref{2.4}). For this purpose we need several lemmas. \begin{lem}\label{l4.1} Suppose that $g$ is a K\"ahler metric on $X$ with constant holomorphic curvature $H$, and let $\Omega$ denote the curvature form of $g$. Then \begin{enumerate} \item With respect to local holomorphic coordinates $(z_1,\ldots,z_n)$, $\Omega$ is given by \begin{equation*} \Omega^j_k=-iH\delta_{jk}\omega+\frac{H}{2}\sum^n_{m=1} g_{k{\overline m}} dz_j\wedge d\overline z_m. \end{equation*} \item The Chern classes of $(X,g)$ are represented by \begin{equation*} c_j(\Omega)= \binom{n+1}{j} \left( -\frac{H}{2\pi}\omega\right)^j. \end{equation*} \end{enumerate} \end{lem} \begin{pf} 1. The claimed formula for $\Omega^j_k$ is an immediate consequence of the following formula for the Riemann curvature tensor $R$ on a manifold with constant holomorphic curvature $H$ \cite[(6.1.1)]{Go}: \begin{equation*} R_{j{\overline k}l{\overline m}}=\frac{H}{2}\left( g_{j{\overline k}} g_{l{\overline m}}+g_{j{\overline m}}g_{l{\overline k}}\right) . \end{equation*} 2. We recall that a representative of the $j$th Chern class is given in terms of the curvature form $\Omega$ by means of the formula \begin{equation*} c_j(\Omega)=\left(\frac{1}{2\pi i}\right)^j\frac{1}{j!} \sum_{\begin{Sb} k_1,\ldots,k_j,l_1,\ldots,l_j\\ \in\{1,\ldots,n\} \end{Sb}} \delta^{l_1\cdots l_j}_{k_1\cdots k_j}\Omega^{k_1}_{l_1}\wedge\cdots\wedge\Omega^{k_j}_{l_j} \end{equation*} \cite[(6.12.1)]{Go}. Inserting the expression for $\Omega^j_k$, given in 1., a straightforward computation leads to \begin{equation*} c_j(\Omega)= \left(\frac{H}{2\pi i}\right)^j\sum^j_{p=0} \binom{n-j+p}{p} \left( -i\omega\right)^j= \binom{n+1}{j}\left( -\frac{H}{2\pi}\omega\right)^j. \end{equation*} \end{pf} Let $U=U_0$, where $U_t$ is defined by (\ref{2.1}). \begin{lem}\label{l4.2} Let $(X,g)$ be the K\"ahler manifold with constant holomorphic curvature $H$. Let $\varphi \in C^\infty(X)$ and let $\omega_u=\omega+iu\partial{\overline\partial}\varphi,\quad |u|<\epsilon$. Then for all $j\ge1$, \begin{equation*} \mbox{Tr} (\Omega^j U)=-\frac{\Delta \varphi}{2}(-iH\omega)^j+ \partial{\overline\partial}\left( \frac{H}{2}\varphi(-iH\omega)^{j-1}\right). \end{equation*} \end{lem} \begin{pf} Choose local holomorphic coordinates $(z_1,\ldots,z_n)$. Then we claim that for all $j\ge1$: \begin{equation}\label{4.2} \begin{split} (\Omega^j U)^p_q= & \sum^n_{k=1}\left\{ (-i H\omega)^jg^{p{\overline k}} \varphi_{q{\overline k}}\right.\\ & + (-iH\omega)^{j-1}\frac{H}{2}\left. \varphi_{q{\overline k}}dz_p\wedge d{\overline z}_k\right\}, \end{split} \end{equation} where $\varphi_{q{\overline k}}=\partial^2/\partial z_q \partial {\overline z}_k\varphi$. To prove (\ref{4.2}), we proceed by induction on $j$. If $j=1$, then using Lemma \ref{l4.1}, we get \begin{equation*} \begin{split} (\Omega U)^p_q & =\sum^n_{l=1}\Omega^p_l U^l_q= \sum_{l,k}\left( -iH\delta_{pl}\omega+\frac{H}{2}\sum^n_{m=1} g_{l{\overline m}}dz_p\wedge dz_{\overline m}\right) g^{l{\overline k}}\varphi_{q{\overline k}}\\ & =\sum^n_{k=1}\left( -iH\omega g^{p{\overline k}}\varphi_{q{\overline k}}+ \frac{H}{2} \varphi_{q{\overline k}}dz_p\wedge{\overline z}_k\right). \end{split} \end{equation*} Suppose that (\ref{4.2}) holds for $j$. Then using Lemma \ref{l4.1} combined with the induction hypothesis, we obtain \begin{equation*} \begin{split} (\Omega^{j+1}U)^p_q& =\sum^n_{l=1}\Omega^p_l(\Omega^jU)^l_q = \sum_{kl}\left( -iH\delta_{pl}\omega+\frac{H}{2}\sum^n_{m=1} g_{l{\overline m}}dz_p \wedge d{\overline z}_m\right)\\ &\wedge\left( (-iH\omega)^jg^{l{\overline k}}\varphi_{q{\overline k}}+(-iH\omega)^{j-1}\frac{H}{2} \varphi_{q{\overline k}} dz_l\wedge d{\overline z}_k\right)\\ &=\sum^n_{k=1}\left\{ (-iH\omega)^{j+1}g^{p{\overline k}} \varphi_{q{\overline k}}+(-iH\omega)^j\frac{H}{2}\varphi_{q{\overline k}}dz_p\wedge d{\overline z}_k\right\}. \end{split} \end{equation*} This proves (\ref{4.2}). Thus in local holomorphic coordinates, we get \begin{equation*} \begin{split} \mbox{Tr}(\Omega^jU)& =(-iH\omega)^j\sum^n_{p,k=1}g^{p{\overline k}} \frac{\partial^2\varphi}{\partial z_p\partial{\overline z}_k}+ \frac{H}{2}(-iH\omega)^{j-1} \sum^n_{p,k=1} \frac{\partial^2\varphi}{\partial z_p\partial{\overline z}_k} dz_p\wedge d{\overline z}_k\\ & = - \frac{\Delta\varphi}{2}(-i H\omega)^j+\frac{H}{2}(-iH\omega)^{j-1} \partial{\overline\partial}\varphi. \end{split} \end{equation*} Since $\omega$ is closed, this implies the lemma. \end{pf} \begin{thm}\label{th4.3} Let $g_{KE}\in\K_\omega$ be a metric of constant holomorphic curvature. Then $g_{KE}$ is a critical point of $\cQ\colon\K_\omega\to\R$. \end{thm} \begin{pf} By (\ref{1.6}) it suffices to prove that \begin{equation*} \frac{\delta}{\delta \varphi }\log{\mathcal Q}(X,g_{KE})=0 \end{equation*} for all $\varphi\in C^\infty(X)$. Let $\varphi\in C^\infty(X)$ and let $g_u$, $|u|<\varepsilon$, be the corresponding variation of $g_{KE}$ in the $\varphi$-direction. We use the variational formula (\ref{2.7}) and apply Lemma \ref{l4.1} and Lemma \ref{l4.2} to the integrant. Then it follows that there exist $A,B\in \C$ such that \begin{equation*} \begin{split} \frac{\delta}{\delta\varphi} \log{\mathcal Q}(X,g_{KE})& = \int_X\left\{ A\Delta\varphi\omega^n+B\partial{\overline\partial}(\varphi\omega^{n-1}) \right\}\\ & = A\int_X\Delta\varphi\omega^n+B\int_X\partial {\overline\partial}(\varphi\omega^{n-1}). \end{split} \end{equation*} Hence, we get $$\frac{\delta}{\delta\varphi} \log{\mathcal Q}(X,g_{KE})=0.$$ This concludes the proof. \end{pf} \noindent In the case of a complex torus, Theorem 0.1 can be restated in terms of the analytic torsion. Namely, we have \begin{cor}\label{c4.4} Let $X=\Gamma\backslash \C^n$ be a complex torus and let $g_{KE}$ be the flat metric on $X$. Then \begin{equation*} \frac{\delta}{\delta\varphi} T_0(X,g_{KE})=0 \end{equation*} for all $\varphi\in C^\infty(X)$. \end{cor} \begin{pf} Let $\varphi\in C^\infty(X)$ and let $g_u,\; |u|<\epsilon$, be the variation of $g_{KE}$ in the $\varphi$-direction. Let $\alpha=\alpha_0$ denote the operator (\ref{2.1}) at $u=0$ where $*_u$ is the Hodge star operator with respect to $g_u$. Let $P_{0,q}, \; 0\le q\le n$, denote the harmonic projections with respect to $g_{KE}$. By Theorem \ref{th4.3} and (\ref{2.5}) it is sufficient to prove that \begin{equation*} \sum^n_{q=0 } (-1)^q Tr(\alpha P_{0,q})=0 \end{equation*} for all $\varphi\in C^\infty(X)$. Let $(z_1,\ldots,z_n)$ denote the standard coordinates on $X$ induced from $\C^n$. Recall that an orthonormal basis for the space $\H^{0,q}(X)$ of harmonic $(0,q)$-forms is given by \begin{equation*} \left\{ (\mbox{vol}(X))^{-1/2} d{\overline z}^B \mid B\subset \{ 1,\ldots,n\},\ |B|=q\right\}. \end{equation*} Using the definition of the Hodge star operator and the assumption that $(g_{KE})_{i{\overline j}}=\delta_{ij},$ it follows that \begin{equation*} \begin{split} \overline{ Tr(\alpha P_{0,q})} & = c\sum_{|B|=q}\int \frac{\partial}{\partial u} \left\{ \det \left(( g^{a{\overline b}}_u)_{a,b\in B}\right)\det\left((g_u)_{i{\overline j}}\right)\right\}\Big|_{u=0} \;dv_0\\ & = c \sum_{|B|=q} \int_X\left\{ \sum_{b\in B} {\dot g}^{b{\overline b}}- \frac{\Delta \varphi}{2}\right\}\;dv_0 \\ &= c_1\int_X\Delta\varphi\;dv_0=0. \end{split} \end{equation*} \end{pf} \begin{remark} If $\K_\omega$ contains a metric $g_{KE}$ of constant holomorphic curvature, then we have seen that $g_{KE}$ is a critical point of $\log{\mathcal Q}(X,\tilde g)$ for $\tilde g\in{\mathcal K}_{\omega}$. We do not know, however, wether or not $g_{KE}$ is the unique critical point of $\log{\mathcal Q}$ on ${\mathcal K}_{\omega}$. Also, we do not know anything about the nature of the critical point. It would be interesting to see wether or not $g_{KE}$ is a maximum or a minimum of $\log{\mathcal Q}(X,\tilde g)$. \end{remark} \section[K3 surfaces]{K3 surfaces} \setcounter{equation}{0} Another class of examples are $K3$ surfaces. Recall that a $K3$ surface is a compact, connected complex analytic surface $X$ that is regular, meaning $h^1(X,{\mathcal O})=0$, and its canonical bundle $K=\Lambda^{2,0}(T^*X)$ is trivial. By Siu \cite{Si}, every $K3$ surface admits a K\"ahler metric. Furthermore, by Yau \cite{Y}, every K\"ahler class $[\omega]$ of $X$ contains a unique K\"ahler-Einstein metric $g_{KE}$. This metric should be a natural candidate for a critical point of $\cQ$ and we shall now investigate under what conditions this is the case. \begin{prop}\label{prop5.1} Let $X$ be a $K3$ surface and let $[\omega]$ be a K\"ahler class of $X$. Then the unique K\"ahler-Einstein metric $g_{KE}\in\K_\omega$ is a critical point of $\cQ:\K_\omega\to\R$ if and only if $\Delta c_2(\Omega_{KE})=0$, where $\Omega_{KE}$ is the curvature form of $g_{KE}$. \end{prop} \begin{pf} Since $c_1(X)=0$, the K\"ahler-Einstein metric $g_{KE}$ is Ricci flat. Therefore we have $c_1(\Omega_{KE})=0$. Then it follows from the variational formula (\ref{2.10}) that $$\frac{\delta}{\delta\varphi}\log\cQ(g_{KE})=\frac{1}{96}\int_X\Delta\varphi c_2(\Omega_{KE})=\frac{1}{96}\int_X\varphi\Delta c_2(\Omega_{KE})$$ for all $\varphi\in C^\infty(X)$. Hence $g_{KE}$ is a critical point of $\cQ$ if and only if $\Delta c_2(\Omega_{KE})=0$. \end{pf} \noindent The condition $\Delta c_2(\Omega_{KE})=0$ can be reformulated in terms of the curvature tensor. \begin{lem}\label{lem5.2} Let $g$ be a Ricci flat metric on $X$. Let $R$ be the curvature tensor of $g$. Then we have $$c_2(\Omega)=\frac{1}{4\pi^2} |R|^2\omega^2,$$ where $| R|$ is the pointwise norm of $R$. \end{lem} \begin{pf} Let $W$ denote the Weyl curvature tensor of $g$. Since $g$ is Ricci flat, it follows from \cite[(1.116)]{Be} that $R=W$. Furthermore, let $L:\Lambda^{p,q}(T^*X)\to\Lambda^{p+1,q+1}(T^*X)$ be the operator $L(\eta)=\omega\wedge \eta$ and let $\Lambda$ be its adjoint. Then it follows from equation (2.80) in \cite{Be} that $$\Lambda^2(c_2(\Omega))=\frac{1}{4\pi^2}|W|^2.$$ This implies the lemma. \end{pf} Combining Proposition \ref{prop5.1} and Lemma \ref{lem5.2}, we obtain \begin{cor}\label{cor5.3} The K\"ahler-Einstein metric $g_{KE}\in\K_\omega$ is a critical point of $\cQ$ if and only if $|R(x)|$ is constant. \end{cor} \smallskip \noindent The condition that $|R(x)|$ is constant is very stringent and we do not know if there exists any $K3$ surface which admits a K\"ahler-Einstein metric satisfying this condition. However, by a result of S. Kobayashi \cite{Ko}, one knows that on certain $K3$ surfaces there exist K\"ahler-Einstein metrics with curvature concentrated near some divisor. These are examples of K\"ahler-Einstein metrics on K3 surfaces such that $|R|$ is not constant. \section[Algebraic manifolds]{Algebraic manifolds} \setcounter{equation}{0} In this section we introduce a twisted version of our functional $\cQ$ which has a simpler variational formula. For this purpose we now assume that $X$ is a projective algebraic manifold. Note that by the Kodaira embedding theorem \cite{GH} any compact complex manifold with definite Chern class is a projective algebraic manifold. \smallskip \noindent A polarization of $X$ is the choice of an ample line bundle $L\to X$ up to isomorphism. The pair $(X,L)$ is called a polarized algebraic manifold. Since $L$ is ample, a certain power $L^{\otimes k}$, $k\in\N$, defines an embedding $i_L: X\hookrightarrow \C P^n$ \cite{GH}. Let $g$ be the pullback of the Fubini-Study metric on $\C P^n$ and let $[\omega]$ be the K\"ahler class determined by $g$. Then $[\omega]$ is a rational multiple of $c_1(L)$. So we can normalize $g$ such that $[\omega]=c_1(L)$. This implies that $[\omega]\in H^2(X,\Z)\cap H^{1,1}(X,\R)$. We fix a hermitian metric $h$ on $L$ such that the curvature $\Theta_h$ satisfies $\Theta_h=-2\pi i\omega$. Such metrics $h$ exist \cite[pp.163,191]{GH} and, up to multiplication by a positive real number, $h$ is uniquely determined by $\omega$. Given $\varphi\in C^\infty(X,\R)$, put \begin{equation*} h(\varphi)=e^{-2\pi\varphi}h. \end{equation*} Then $h(\varphi)$ is a hermitian metric on $L$ whose curvature is given by \begin{equation*} \Theta_{h(\varphi)}=\Theta_h+2\pi\partial{\overline\partial}\varphi=-2\pi i(\omega +i\partial{\overline{\partial}}\varphi). \end{equation*} Thus, if $\varphi\in{\mathcal H}_\omega$, we have \begin{equation}\label{6.1} \Theta_{h(\varphi)}=-2\pi i\omega(\varphi). \end{equation} In this section we consider the analytic torsion with coefficients in a certain virtual vector bundle $\E$ associated to $L$. For its definition we introduce the following numbers \begin{equation*} \kappa_1=\int_Xc_1(\Omega)\wedge \omega^{n-1} \;\;\mbox{and}\;\; \kappa_2=\int_X\omega^n. \end{equation*} Since $[\omega]\in H^2(X,\Z)$, it follows that $\kappa_1$ and $\kappa_2$ are integers. Then by (\ref{1.2}), we have $$s_0=4n\pi\frac{\kappa_1}{\kappa_2}.$$ Now we define the virtual vector bundle $\E$ over $X$ to be \begin{equation*} \E=\bigoplus^{4\kappa_2(n+1)} (L-L^{-1})^{\otimes n}\oplus \bigoplus^{-\kappa_1n}(L-L^{-1})^{\otimes(n+1)}. \end{equation*} If $\widetilde h$ is any hermitian metric on $L$, we denote by ${\widetilde h}^{\E}$ the induced hermitian metric on $\E$. Using this notation we introduce the following functional on $\H_\omega$: \begin{equation}\label{6.3} \cQ_L(X,\varphi)={\mathcal Q}(X,\E,g(\varphi),h(\varphi))^{(-1)^n},\quad \varphi\in\H_\omega. \end{equation} Since $X$ is fixed, we set $\cQ_L(\varphi)=\cQ_L(X,\varphi)$. To begin with we compute the variation of $\cQ_L(\varphi)$. \begin{thm}\label{th6.1} Let $\varphi_u,\;u\in[a,b]$, be a smooth family of functions in ${\mathcal H}_\omega$. Let $\omega_u=\omega+i\partial\overline\partial\varphi_u$ and let $s(u)$ be the scalar curvature of the metric $g_u=g(\varphi_u)$. Then \begin{equation*} \frac{\partial}{\partial u}\log \cQ_L(\varphi_u)=c_n\int_X{\dot\varphi}_u (s(u)-s_0)\omega^n_u, \end{equation*} where $$c_n=\kappa_2(n+1)2^{n-1}.$$ \end{thm} \begin{pf} To compute the variation, we apply formula (\ref{2.3}). Let $F^{\E}_u$ denote the curvature form of the metric $h^{\E}_u$, induced on $\E$ by $h(\varphi_u)$, and let $V_u=(h^\E_u)^{-1} \frac{d}{du}(h^{\E}_u)$. Then by (\ref{2.3}) we have \begin{equation}\label{6.4} \begin{split} \frac{\partial}{\partial u}\log \cQ_L(\varphi_u)& =\frac{(-1)^n}{2} \left(\frac{1}{2\pi i}\right)^n\int_X\left[ \frac{\partial}{\partial b}\operatorname{Td}(-\Omega_u-bU_u)\Big|_{b=0}\operatorname{ch}(\E,h^\E_u)\right.\\ & +\left.\operatorname{Td}(-\Omega_u)\frac{\partial}{\partial b} \operatorname{ch} (-F^\E_u-b V_u)\Big|_{b=0}\right]. \end{split} \end{equation} \noindent We start with the computation of the first term. Using that $\Theta_u=\Theta_{h_{u}}=-2\pi i \omega_u$, it follows that \begin{equation*} \begin{split} \operatorname{ch}(\E,h^\E_u)&=4\kappa_2(n+1)(e^{2\pi i\omega_u}-e^{-2\pi i\omega_u})^n- \kappa_1 n(e^{2\pi i\omega_u} -e^{-2\pi i\omega_u})^{n+1}\\ & =4\kappa_2(n+1)(4\pi i\omega_u)^n. \end{split} \end{equation*} Hence, by Lemma \ref{l2.1} we obtain \begin{equation}\label{6.5} \begin{split} \int_X\biggl[ \frac{\partial}{\partial b}\operatorname{Td}(&-\Omega_u-bU_u)\Big|_{b=0} \operatorname{ch}(\E,h^\E_u)\biggr]^{(n)}\\ &=2\kappa_2(n+1)(4\pi i)^n\int_X\frac{\partial}{\partial b}c_1(-\Omega_u-bU_u)\Big|_{b=0}\omega_u^n\\ & =\kappa_2(n+1)(4\pi i)^n\int_X(\Delta_u\varphi_u)\omega^n_u=0. \end{split} \end{equation} As for the second term, we have \begin{equation*} \begin{split} \frac{\partial}{\partial b}\operatorname{ch} \bigl( &-F^\E_u-bV_u\bigr) \Big|_{b=0}\\ &=(n+1)\left\{ 4n\kappa_2 \operatorname{ch} (L_u^{-1}-L_u)^{n-1}-n\kappa_1 \operatorname{ch}(L_u^{-1}-L_u)^n\right\}\\ & \hskip20pt\cdot \frac{\partial}{\partial b}\left\{ \operatorname{ch}(2\pi i\omega_u+b 2\pi {\dot\varphi}_u)-\operatorname{ch}(-2\pi i\omega_u-b2\pi\dot\varphi_u)\right\}\Big|_{b=0}\\ & =2\pi (n+1)\left\{ 4n\kappa_2(e^{-2\pi i\omega_u} -e^{2\pi i\omega_u})^{n-1}\right.\\ &\hskip43pt\left.-n\kappa_1(e^{-2\pi i\omega_u}-e^{2\pi i\omega_u})^n\right\}\dot\varphi_u (e^{2\pi i\omega_u}+ e^{-2\pi i\omega_u})\\ & =4\pi (n+1)\kappa_2(-4\pi i)^n\left\{-\frac{n}{\pi i} \omega^{n-1}_u-n\frac{\kappa_1}{\kappa_2}\omega^n_u\right\} \dot\varphi_u.\\ \end{split} \end{equation*} \noindent By (\ref{1.1}), we have $c_1(\Omega_u)=\rho/2\pi$. Hence, using (\ref{1.2}), we get \begin{equation*} \begin{split} \Biggl[ \operatorname{Td}(-&\Omega_u)\frac{\partial}{\partial b}\operatorname{ch}(-F^\E_u-bV_u)\Big|_{b=0}\Biggr]^{(n)}\\ & = 4\pi (n+1)\kappa_2(-4\pi i)^n\left(nc_1(\Omega_u)\wedge \omega^{n-1}_u-n\frac{\kappa_1}{\kappa_2}\omega^n_u\right)\dot\varphi_u\\ &=2^{n}(n+1)\kappa_2(2\pi i)^n(-1)^n\dot\varphi_u(s(u)-s_0)\omega_u^n. \end{split} \end{equation*} \noindent Together with (\ref{6.4}) and (\ref{6.5}) this leads to \begin{equation*} \frac{\partial}{\partial u}\log \cQ_L(\varphi_u)=c_n\int_X \dot\varphi_u (s(u)-s_0) \omega_u^n. \end{equation*} \end{pf} \begin{remark} There are other possible choices for $\E$ which give essentially the same variational formula. Tian \cite{Ti1} used in a different context a virtual bundle of the form $$\bigoplus^a\left((K_X^{-1}-K_X)\otimes(L-L^{-1})^{n}\right)\oplus \bigoplus^b(L-L^{-1})^{n+1}.$$ Up to a constant, it gives the same variational formula. \end{remark} \noindent Let $\varphi_1,\varphi_2\in{\mathcal H}_\omega$. Let $\{\varphi_u\mid a\le u\le b\} $ be a piecewise smooth path in ${\mathcal H}_\omega$ such that $\varphi_a=\varphi_1$ and $\varphi_b=\varphi_2$. Then it follows from Theorem \ref{th6.1} that \begin{equation}\label{6.6} \log \left( \frac{\cQ_L(\varphi_1)}{\cQ_L(\varphi_2)}\right)=-c_n\int^b_a \left\{ \int_X \dot\varphi_u (s(u)-s_0)\omega^n_u\right\} \; dt. \end{equation} \noindent The right hand side of (\ref{6.6}) is equal to $c_n$ times the functional $M(\varphi_1,\varphi_2)$ introduced by Mabuchi \cite[(2.2.2)]{Ma1}. It is defined for any compact K\"ahler manifold $X$. For a projective algebraic manifold $X$, (\ref{6.6}) implies that $M(\varphi_1,\varphi_2)$ is independent of the path $\{ \varphi_u\mid a\le u\le b\}$ in ${\mathcal H}_\omega$ connecting $\varphi_1$ and $\varphi_2$. This was proved by Mabuchi \cite[Theorem 2.4]{Ma1} in general, using different methods. \smallskip \noindent Let $\varphi\in{\mathcal H}_\omega$ and $c\in\R$. Set $\varphi_u=\varphi+uc$, $u\in[0,1]$. Then by (\ref{6.6}) and (\ref{1.2}) we get \begin{equation*} \begin{split} \log\left( \frac{\cQ_L(\varphi+c)}{\cQ_L(\varphi)}\right)&=c_nc\int^1_0\left\{ \int_X(s(u)-s_0)\omega^n_u\right\}\;du\\ & = \frac{c_nc}{n!} \int^1_0 \left\{ \int_Xs(u)dv_g-s_0\mbox{Vol}(X)\right\} \;du=0. \end{split} \end{equation*} \noindent Hence for all $\varphi\in{\mathcal H}_\omega$ and $c\in\R$ we have \begin{equation*} \cQ_L(\varphi+c)=\cQ_L(\varphi). \end{equation*} Thus by (\ref{1.6}), $\cQ_L:{\mathcal H}_\omega\to \R$ factors through ${\mathcal K}_\omega$. The induced functional on ${\mathcal K}_\omega$ will be denoted by the same letter: \begin{equation*} \tilde g\in{\mathcal K}_{\omega}\longmapsto \cQ_L(\tilde g)\in\R. \end{equation*} By (\ref{6.6}), the map $\mu\colon{\mathcal K}_{\omega_{0}}\to\R$ defined by \begin{equation}\label{6.7} \mu(\tilde g)= c_n^{-1}\log \frac{\cQ_L(g)}{\cQ_L(\tilde g)} \end{equation} coincides with the $K$-energy map defined by Mabuchi \cite[Section3]{Ma1}. \noindent Let $g_u\in\K_\omega$, $|u|< \epsilon$, be a smooth path. Then by (\ref{1.6}) there exists a smooth path $\varphi_u\in\H_\omega$ such that $g_u=g(\varphi_u)$. It follows from Theorem \ref{6.1} that \begin{equation*} \frac{\partial}{\partial u}\log\cQ_L(g_u)\Big|_{u=0}=c_n \int_X\dot\varphi_0(s_{g_0}-s_0)\omega_0^n. \end{equation*} In particular, if $g_u$, $|u|<\epsilon$, is the variation of $g$ in the direction of some $\varphi\in C^\infty(X)$, then this formula implies \begin{equation*} \frac{\delta}{\delta\varphi}\log\cQ_L(g)=c_n \int_X\varphi(s_g-s_0)\omega^n. \end{equation*} This proves the following theorem. \begin{thm}\label{th6.2} A K\"ahler metric $g_0\in{\mathcal K}_{\omega}$ is a critical point of ${\mathcal Q}_L\colon\K_{\omega}\to\R$ if and only if the scalar curvature $s_{g_0}$ of $g_0$ is constant. In this case $s_{g_0}=s_0$. \end{thm} \smallskip \noindent This holds for the K-energy in general \cite{Ma1}. \smallskip \noindent In \cite{Ca}, Calabi introduced the notion of extremal K\"ahler metrics where a K\"ahler metric $\tilde g\in{\mathcal K}_{\omega}$ is called extremal if $\tilde g$ is a critical point of the functional \begin{equation*} S(\widetilde g)=\int_X s^2_{\widetilde g}\; dv_{\widetilde g},\quad {\widetilde g}\in {\mathcal K}_{\omega}. \end{equation*} Any metric $\tilde g$ of constant scalar curvature is extremal, but as shown by Calabi \cite{Ca}, the converse is not true, i.e., there are extremal K\"ahler metrics with nonconstant scalar curvature. However, if $X$ is non-uniruled, then the extremal K\"ahler metrics are precisely the metrics with constant scalar curvature. \smallskip \noindent Now we collect some further information about the functional $\cQ_L$. By Theorem \ref{th6.2}, the critical points of $\cQ_L$ are the metrics of constant scalar curvature. If $\K_\omega$ contains a K\"ahler-Einstein metric, then the subset $\K_E\subset \K_\omega$ of all K\"ahler-Einstein metrics coincides with the set of critical points of $\cQ_L$. In particular, if $c_1(X)\le0$, then by \cite{A}, \cite{Y}, $\cQ_L$ has critical points. On the other hand, the Futaki invariant \cite{Fu}, \cite{Ca2} obstructs the existence of constant scalar curvature metrics in certain K\"ahler classes. Therefore, critical points do not always exist. The K-energy map $\mu\colon\K_\omega\to\R$ has been studied by Mabuchi and Bando \cite{Ma1}, \cite{Ma2}, \cite{B}, \cite{BM}, mainly in connection with the Futaki obstruction to the existence of K\"ahler-Einstein metrics on compact complex manifolds with $c_1(X)>0$. Mabuchi \cite{Ma2} defined a natural Riemannian structure on $\K_\omega$, i.e., $\K_\omega $ can be equipped canonically with the structure of an infinite-dimensional Riemannian manifold. The tangent space $T_{\tilde\omega}\K_\omega$ of $\K_\omega$ at $\tilde\omega\in\K_\omega$ can be identified with $$\left\{\eta\in C^\infty(X,\R)\mid \int_X\eta\tilde\omega^n=0\right\}$$ and the Riemannian structure is given by $$\langle\eta_1,\eta_2\rangle=\frac{1}{\text{Vol(X)}} \int_X\eta_1\eta_2\frac{\tilde\omega^n}{n!},\quad\eta_1,\eta_2\in T_{\tilde\omega}\K_\omega.$$ One of the main results of \cite[Theorem 5.3]{Ma2} states that $\mu\colon\K_\omega\to\R$ is a convex function, meaning that $\text{Hess}\:\mu$ is positive semidefinite. Therefore, by (\ref{6.7}) the same holds for $\log\cQ_L$. This resembles the situation in the Riemann surface case. If $\K_\omega$ contains a K\"ahler-Einstein metric, then by Theorem \ref{0.1}, $\cQ_L$ is bounded from above. In general, we do not know anything about boundedness of $\cQ_L$. Finally, we note that by Theorem \ref{6.1} the gradient flow of $\log\cQ_L$ is given by $$\frac{\partial\varphi }{\partial t}(t)=s(t)-s_0,$$ where $s(t)$ is the scalar curvature of the metric $g(t)=g(\varphi(t))$. Using that $$s=2\sum_{kl}g^{k\overline l}r_{k\overline l}\quad\text{and}\quad r_{k\overline l}=-\frac{\partial^2}{\partial z_k\partial\overline z_l} \log\det(g_{\alpha\overline \beta}),$$ we get the following fourth order nonlinear parabolic equation \begin{equation}\label{6.8} \frac{\partial\varphi}{\partial t}=-2\sum_{k,l}g(t)^{k\overline l}\frac{\partial^2}{\partial z_k\partial\overline{z_l}}\left(\log\det\left(g_{\alpha\overline \beta} +\frac{\partial^2\varphi}{\partial z_\alpha\partial\overline{z_\beta}}\right)\right)-s_0. \end{equation} Differentiating this equation with respect to $t$, we get $$\frac{\partial}{\partial t}\left(\frac{\partial\varphi}{\partial t}\right)=-2\Delta_t^2\left(\frac{\partial\varphi}{\partial t}\right)+\sum_{k,l}r(t)^{k\overline l}\frac{\partial^2}{\partial z_k\partial\overline{z}_l}\left(\frac{\partial \varphi}{\partial t}\right),$$ where $r(t)$ is the Ricci tensor of the metric $g(t)$. So, from standard theory we know that the solution of the initial value problem for (\ref{6.8}) exists for short time. The crucial question is to prove existence for all time. If $\varphi(t)$ is any solution of (\ref{6.8}), then by Theorem \ref{6.1} we get $$\frac{\partial}{\partial t}\log\cQ_L(g(\varphi(t))= c_n\int_X(s(t)-s_0)^2\omega_t^n\ge0.$$ Thus $\log\cQ_L$ is increasing along the gradient flow. \section[K\"ahler-Einstein metrics]{K\"ahler-Einstein metrics} \setcounter{equation}{0} \noindent By Theorem \ref{th6.2}, the critical points of $\cQ_L$ are exactly the metrics of constant scalar curvature. In general, it seems to be difficult to determine the nature of the critical points. Much more can be said if $[\omega]$ contains a K\"ahler-Einstein metric $\omega_{KE}$. Recall that $\omega$ is said to be K\"ahler-Einstein if the Ricci form $\rho_\omega$ is proportional to the K\"ahler form \begin{equation}\label{7.1} \rho_\omega=\lambda \omega \end{equation} for some $\lambda\in\R$. Since $2\pi c_1(X)=[\rho_\omega]$, the first Chern class has to satisfy either one of the following conditions \begin{equation*} c_1(X)=0\quad \mbox{or}\quad c_1(X)>0\quad\mbox{or}\quad c_1(X)<0. \end{equation*} If $c_1(X)=0$, it follows from Yau \cite{Y} that each K\"ahler class $[\omega]$ contains a unique K\"ahler-Einstein metric $\omega_{KE}$. If $c_1(X)<0$, i.e., if the canonical line bundle $K_X$ is ample, then it was shown by T. Aubin \cite{A} and S.-T. Yau \cite{Y}, that up to multiplication by a positive scalar, there exists a unique K\"ahler-Einstein metric $g_{KE}$ on $X$ which can be normalized such that $2\pi c_1(X)=-[\omega_{KE}]$. In the case $c_1(X)>0$ there are obstructions for the existence of K\"ahler-Einstein metrics \cite{Fu}, \cite{Ti2} and the existence problem has not been completely settled yet. See \cite{Bo} for a review of the results. First we make the following elementary observation. \begin{lem}\label{l7.1} Let $(X,g)$ be a compact K\"ahler manifold with definite or vanishing first Chern class. Suppose that there is a K\"ahler-Einstein metric $g_{KE}$ on $X$ with $[\omega_{KE}]=[\omega_g]$. Then $g$ has constant scalar curvature iff $g$ is K\"ahler-Einstein. \end{lem} \begin{pf} It follows from (\ref{1.1}) that a K\"ahler-Einstein metric has constant scalar curvature. Now assume that $g$ has constant scalar curvature $s_g$. Then $s_g=s_0$. Let $g_{KE}$ be a K\"ahler-Einstein metric with $[\omega_{KE}]=[\omega]$. By (\ref{1.1}), we have \begin{equation*} \rho_{KE}=\frac{s_0}{2n}\omega_{KE}. \end{equation*} Since $[\rho_{KE}]=[\rho]$, we get $\frac{s_0}{2n}[\omega]=[\rho]$. Hence there exists $\varphi \in C^\infty(X,\R)$ such that \begin{equation*} \rho=\frac{s_0}{2n}\omega+i\partial{\overline\partial}\varphi. \end{equation*} Taking the trace of both sides of this equation yields \begin{equation*} \frac{s_0}{2}-\frac{1}{2}\Delta\varphi=\text{tr}(r)=\frac{s_g}{2} =\frac{s_0}{2}. \end{equation*} This implies that $\varphi$ is constant. Therefore, we get $\rho=\frac{s_0}{2n}\omega$. \end{pf} \begin{cor}\label{cor7.2} Let $\K_{KE}\subset\K_\omega$ denote the set of all K\"ahler-Einstein metrics contained in $\K_\omega$. Suppose that $\K_{KE}\not=\emptyset$. Then $\K_{KE}$ is precisely the set of critical points of $\cQ_L$. \end{cor} \smallskip \noindent {\bf Proof of Theorem 0.1.} First assume that $c_1(X)\le 0$. Since by our assumption, the set of K\"ahler-Einstein metrics which are contained in $\K_\omega$ is nonempty, it follows from \cite{A} and \cite{Y} that $\K_\omega$ contains a unique K\"ahler-Einstein metric $g_{KE}$, and if $c_1(X)<0$, then $\frac{1}{2\pi}[\omega_{KE}]$ represents $-c_1(X)$. Now it follows from Corollary \ref{cor7.2} that $g_{KE}$ is the unique critical point of $\cQ_L:\K_{\omega}\to\R$. To prove that $g_{KE}$ is the absolut maximum of $\cQ_L$, we employ the complex analogue of the Ricci flow. We normalize $g_{KE}$ such that \begin{equation*} \rho_{KE}=\frac{s_0}{2n}\omega_{KE}. \end{equation*} Let $g\in\K_{\omega}$. By H.-D. Cao \cite{C}, the evolution equation \begin{equation}\label{7.2} \frac{\partial {\widetilde g}_{i{\overline j}}}{\partial t} =-{\widetilde r}_{i{\overline j}}+\frac{s_0}{2n}{\widetilde g}_{i{\overline j}},\quad {\widetilde g}_{i{\overline j}}(0)=g_{i{\overline j}},\quad i,j=1,\ldots,n, \end{equation} has a unique solution ${\widetilde g}(t)$ which exists for all $t\ge 0$ and satisfies \begin{equation}\label{7.3} \lim_{t\to \infty} {\widetilde g}(t)=g_{KE} \end{equation} in the $C^\infty$-topology. Moreover, there exists $u\in C^\infty(\R^+\times X,\R)$ with \begin{equation}\label{7.4} {\widetilde g}_{i{\overline j}}(t,z)=g_{i{\overline j}}(z)+ \frac{\partial^2u(t,z)}{\partial z_i\partial{\overline z}_j}. \end{equation} Let $\Delta_t$ be the Laplace operator with respect to ${\widetilde g}(t)$. By (\ref{7.2}) and (\ref{7.4}) we obtain \begin{equation*} \begin{split} \frac{1}{2}\Delta_t({\dot u}(t))& =-\sum_{i,j}{\widetilde g}^{i{\overline j}}(t) \frac{\partial^2{\dot u}(t)}{\partial z_j\partial {\overline z}_j}=-\sum_{i,j}{\widetilde g}^{i{\overline j}}(t)\frac{\partial {\widetilde g }_{i{\overline j}}}{\partial t}(t)\\ &= \sum_{i,j}{\widetilde g}^{i{\overline j}}(t) \left( \widetilde r_{i{\overline j}}(t) -\frac{s_0}{2n} {\widetilde g}_{i{\overline j}}(t)\right)=\frac{{\widetilde s}(t)}{2}- \frac{s_0}{2}. \end{split} \end{equation*} Thus $u(t)$ satisfies \begin{equation}\label{7.5} \Delta_t(\dot u(t))=\widetilde s(t)-s_0. \end{equation} Together with Theorem \ref{th6.1} we get \begin{equation}\label{7.6} \begin{split} \frac{\partial}{\partial t}\log \cQ({\widetilde g}_t) & = c_n\int_X{\dot u}(t)({\widetilde s}(t)-s_0)\omega^n_t\\ & = c_n\int_X{\dot u}(t)\Delta_t({\dot u}(t))\omega^n_t\\ &= \frac{c_n}{n!}\int_X|\nabla \dot u(t)|_t^2\;dv_t. \end{split} \end{equation} Hence, by the definition of $c_n$, we obtain \begin{equation}\label{7.7} \frac{\partial}{\partial t}\log\cQ(\widetilde g_t)\ge0. \end{equation} Suppose that $\widetilde s(0)$ is not constant. Then by (\ref{7.5}) there exists $\varepsilon>0$ such that for all $t\le\varepsilon$, $\dot u(t,z)$ is not constant as a function of $z$. By (\ref{7.6}) it follows that \begin{equation}\label{7.8} \frac{\partial}{\partial t}\log\cQ(\widetilde g_t)>0\quad\text{for all}\quad t\le\varepsilon. \end{equation} By \cite[Proposition 2.2]{C}, ${\dot u}(t)$ converges to a constant in the $C^\infty$-topology as $t\to\infty$. Also ${\widetilde g}_{i{\overline j}}(t)$ converges to $(g_{KE})_{i{\overline j}}$ in the $C^\infty$-topology as $t\to\infty$ \cite[Main Theorem]{C}. This implies that there exists $C_1>0$ such that \begin{equation}\label{7.9} \sup_{z\in X}| \Delta_t({\dot u}(t,z))|\le C \end{equation} for all $t\ge 0$. Put \begin{equation*} a(t)=\frac{1}{\mbox{Vol}(X)} \int_X{\dot u}(t) dv_t, \; t\ge 0. \end{equation*} By (2.11) and (2.18) of \cite{C} there exist $C_2,C_3>0$ such that \begin{equation}\label{7.10} \int_X|{\dot u}_t-a(t)|dv_t\le C_2 e^{-C_3t}. \end{equation} Using (\ref{7.9}) and (\ref{7.10}), we obtain \begin{equation}\label{7.11} \left| \int_X{\dot u}_t\Delta({\dot u}_t)dv_t\right|= \left| \int_X({\dot u}_t-a(t))\Delta_t({\dot u}_t)dv_t\right| \le C_4 e^{-C_3t}. \end{equation} Since for $t\to\infty$, $\widetilde g_{i\overline j}(t)$ converges to $(g_{KE})_{i\overline j}$ in the $C^\infty$-topology, it follows that \begin{equation}\label{7.13} \log\cQ_L(g_{KE})=\lim_{t\to\infty} \log \cQ_L({\widetilde g}_t). \end{equation} Combing (\ref{7.6}), (\ref{7.11}) and (\ref{7.13}), it follows that for all $t>0$, we have \begin{equation*} \log \cQ(g_{KE})-\log \cQ({\widetilde g}_t)=c_n\int^\infty_t \frac{\partial}{\partial v}\log \cQ({\widetilde g}_v)dv. \end{equation*} Hence, if $g\not=g_{KE}$, (\ref{7.7}) together with (\ref{7.8}) imply that \begin{equation*} \log \cQ_L(g_{KE})> \log \cQ_L(g). \end{equation*} This completes the proof of the first part of the theorem. \smallskip \noindent Now assume that $c_1(X)>0$. Let $\K_{E}\subset\K_0$ be the subset of K\"ahler-Einstein metrics. By our assumption $\K_{E}\not=\emptyset$. Then it follows from \cite[Theorem 1]{B} that the K-energy map $\mu\colon\K_\omega\to\R$ takes its absolute minimum on $\K_{E}$. Hence, by (\ref{6.7}), $\cQ_L$ attains its maximum on $\K_{E}$. \section[Moduli spaces]{Moduli spaces and Quillen metric} \setcounter{equation}{0} So far, we considered the Quillen norm $\cQ_L$ of a fixed vector $v\in\lambda(\E)$ as a function on $K_\omega$. In this section we investigate the behaviour of the Quillen metric with respect to variations of the complex structure. In \cite{FS}, Fujiki and Schumacher defined the moduli space of extremal K\"ahler metrics on a fixed $C^\infty$ manifold. First we consider the local problem. Let $(\pi:\X\to S,\widetilde\omega)$ be a metrically polarized family of compact K\"ahler manifolds \cite[Definition 3.2]{FS} over a reduced complex space $S$ and let $p\colon F\to \X$ be a holomorphic hermitian vector bundle with metric $h^F$. Let $\lambda(F)$ be the holomorphic line bundle on $S$ associated to $(\det R\pi_*\F)^{-1}$, where $\F$ is the locally free sheaf corresponding to $F$. Given $s\in S$, let $X_s=\pi^{-1}(s)$ and $F_s=F| X_s$. We recall that there exists a canonical isomorphism \begin{equation}\label{8.1} \lambda(F)_s\cong\lambda(F_s)=\bigotimes_{q\ge0}\left(\det H^q(X_s,F_s)\right)^{(-1)^{q+1}} \end{equation} \cite{BGS3}. For each $s\in S$, $\tilde \omega$ defines a K\"ahler metric $g_s$ on $X_s$ and we get a family $\tilde g=\{g_s\}_{s\in S}$ of K\"ahler metrics. Let $h_s^F$ be the hermitian metric on $F_s$ induced by $h^F$. Let $h^{\lambda(F_s)}$ be the Quillen metric on $\lambda(F_s)$ associated to $(g_s,h^F_s)$. Using the isomorphism (\ref{8.1}), $\{h^{\lambda(F_s)}\}_{s\in S}$ defines a metric $h^{\lambda(F)}$ on $\lambda(F)$. It is proved in \cite{BGS3} that the metric $h^{\lambda(F)}$ is smooth. This metric is called the {\it Quillen metric} on $\lambda(F)$ associated to $(\tilde g,h^F)$. If $S$ is a complex manifold, then the curvature of the holomorphic connection on $\lambda(F)$ was computed by \cite[Theorem 1.27]{BGS3}. Namely the first Chern form is given by \begin{equation}\label{8.2} c_1(\lambda(F),h^{\lambda(F)})=-\left[\int_{\X/S}\operatorname{Td}(\X/S,\widetilde g) \operatorname{ch}(F,h^F)\right]^{(2)} \end{equation} where $\operatorname{Td}(\X/S,\widetilde g)$ is the Todd class of the relative tangent bundle. This formula was extended by Fujiki and Schumacher \cite[Theorem 10.1]{FS} to the case where $S$ is a reduced complex space. Now let $(\pi:\X\to S,\widetilde\omega,\L)$ be a metrically polarized family of compact Hodge manifolds \cite[Definition 3.8]{FS}. Thus $(\pi:\X\to S,\L)$ is a polarized family of Hodge manifolds and $(\pi:\X\to S,\widetilde\omega)$ is a metrically polarized family of K\"ahler manifolds such that for each $s\in S$, $\omega_s$ represents $c_1(L_s)$ on $X_s$. Furthermore, there exists a hermitian metric $h$ on $\L$ such that the restriction of $c_1(\L,h)$ to each fibre $X_s$ represents $\omega_s$ \cite[Proposition 3.10]{FS}. Any such metric is called admissible. Let $$\kappa_1=\int_{X_s}c_1(X_s)\wedge\omega_s^{n-1}\quad\text{and}\quad\kappa_2=\int_{X_s}\omega_s^n.$$ These are integers and therefore, they are independent of $s\in S$. Using $\L$, we introduce the family version of the virtual holomorphic bundle used in Section 6. Let $n$ be the fibre dimension and set $$\E=\bigoplus^{4(n+1)\kappa_2}(\L-\L^{-1})^n\oplus\bigoplus^{-n\kappa_1}(\L-\L^{-1})^{n+1}.$$ Let $h$ be an admissible hermitian metric on $\L$ for $(\pi,\widetilde\omega)$. Let $h^\E$ be the induced hermitian metric on $\E$. Then the Chern character of $(\E,h^\E)$ is given by \begin{equation*} \begin{split} \text{ch}(\E,h^\E)=4(n+1)\kappa_2\Biggl\{2^n&c_1(\L,h)^n +\frac{n2^{n-1}}{3}c_1(\L,h)^{n+2}+\cdots\Biggr\}\\ &-n\kappa_1\left\{2^{n+1}c_1(\L,h)^{n+1}+\cdots\right\}. \end{split} \end{equation*} Hence by (\ref{8.2}) we get the following expression for the first Chern form of $(\lambda(\E),h^{\lambda(\E)})$: \begin{equation}\label{8.3} \begin{split} c_1(\lambda(\E),h^{\lambda(\E)})=-\kappa_2(n+1)&2^{n+1}\Biggl(\int_{\X/S} c_1(\L,h)^nc_1(\X/S,\widetilde g)\\ &-\frac{n}{n+1}\frac{\kappa_1}{\kappa_2}\int_{\X/S}c_1(\L,h)^{n+1}\Biggr). \end{split} \end{equation} Assume that the family $(\pi:\X\to S,\widetilde\omega,\L)$ is effective which means that the Kodaira-Spencer map associated to $(\pi,\L)$ is injective \cite[Definition 4.1]{FS}. Then by \cite[Definition 7.2]{FS}, there exists a generalized Weil-Petersson metric $\widehat h_{WP}=\{\widehat h_s\}_{s\in S}$ on $S$. Let $\widehat\omega_{WP}$ denote the corresponding Weil-Petersson form. Using Theorem 7.8 of \cite{FS} together with (\ref{8.3}), one gets the following theorem which is analogous to Theorem 10.3 of \cite{FS}. \begin{thm}\label{th8.1} Let $(\pi:\X\to S,\widetilde\omega,\L)$ be an effective family of extremal compact Hodge manifolds of constant scalar curvature with $S$ being connected, and let $h$ be an admissible hermitian metric on $\L$ for $(\pi,\widetilde\omega)$. Then the first Chern form of the determinant line bundle $(\lambda(\E),h^{\lambda(\E)} )$ is given by $$c_1(\lambda(\E),h^{\lambda(\E)} )=a_n\widehat\omega_{WP}$$ where $a_n=-2^{n+1}\kappa_2(n+1)!$. \end{thm} If we recall the isomorphism (\ref{8.1}) and the construction of the Quillen metric on $\lambda(\E)$, we get an expression of the curvature in terms of the analytic torsion. For each $s\in S$, the fibre $X_s$ is equipped with an extremal Hodge metric $g_s$ and $\L_s$ is equipped with a hermitian metric $h_s$ such that $\Theta_{h_s}=-{2\pi i}\omega_s$. Let $h^{\E_s}$ be the associated hermitian metric in the virtual bundle $$\E_s=\bigoplus^{4(n+1)\kappa_2}(\L_s-\L_s^{-1})^n\oplus\bigoplus^{-n\kappa_1}(\L_s-\L_s^{-1})^{n+1}$$ and put $T_0(X_s,\E_s)=T_0(X_s,\E_s,g_s,h^{\E_s})$. Furthermore, let $\parallel\cdot\parallel_{\lambda(\E_s)}$ denote the metric induced by $(g_s,h^{\E_s})$ on the determinant line $$\lambda(\E_s)=\bigotimes_{q\ge0}\left(\det H^q(X_s,\E_s)\right)^{(-1)^{q+1}}.$$ Now let $\phi:S^\prime\to\lambda(\E)$ be a local holomorphic section of $\lambda(\E)$ over some open subset $S^\prime\subset S$. Using the definition of the Quillen metric in $\lambda(\E)$, it follows from Theorem \ref{th8.1} that \begin{equation}\label{8.4} \partial_s\overline\partial_s\log\left(\parallel\phi(s)\parallel^2_{\lambda(\E_s)}T_0(X_s,\E_s)\right)=a_n\widehat{\omega}_{WP}. \end{equation} In other words, $\log\left(\parallel\phi(s)\parallel^2_{\lambda(\E_s)}T_0(X_s,\E_s)\right)$ is a potential for the generalized Weil-Petersson metric on $S$. \smallskip \noindent This construction can be globalized \cite[\S 11]{FS}. Fix a compact connected $C^\infty$ manifold $X$ and an integral class $\alpha\in H^2(X,\R)$. Let ${\mathfrak M}_{H,e}$ be the moduli space of extremal Hodge manifolds with underlying $C^\infty$ structure $(X,\alpha)$. Then ${\mathfrak M}_{H,e}$ is a complex $V$-manifold and the generalized Weil-Petersson metric is a $V$-form $\omega_{WP}$ on ${\mathfrak M}_{H,e}$. As shown in \cite{FS}, the local determinant line bundles can be patched together to a holomorphic line bundle $F$ on ${\mathfrak M}_{H,e}$ and the Quillen metric on the local bundles patches together to a $V$-metric $h^F$ on $F$. The global version of Theorem \ref{th8.1} implies that the Chern $V$-form of $(F,h^F)$ satisfies \begin{equation}\label{8.5} c_1(F,h^F)=a\omega_{WP} \end{equation} for some integer $a\not=0$. In particular, this implies that the cohomology class $[\omega_{WP}]$ is an integral cohomology class. Hence, any compact analytic subspace of ${\mathfrak M}_{H,e}$ is a projective algebraic manifold. We note that this is well known for Riemann surfaces. Let $X$ be a compact Riemann surface of genus $g\ge2$. Then there exists a discrete cocompact subgroup $\Gamma\subset \text{SL}(2,\R)$ such that $X=\Gamma\backslash H$, where $H$ is the upper half-plane. Let $g$ be the Riemannian metric on $X$ which is induced by the Poincar\'e metric on $H$ and let $\Delta=\overline\partial^*\overline\partial$ be the corresponding Laplace operator. Fix a marking of $X$ and let $\tau$ be the corresponding period matrix of $X$. We regard both $\Delta$ and $\tau$ locally as functions on the moduli space ${\mathfrak M}_g$ of compact Riemann surfaces of genus $g$. Then one has $$\partial_z\overline\partial_z\log\left(\frac{\det\Delta_z} {\det\tau_z}\right)=\frac{i}{6\pi}\omega_{WP},$$ where $\omega_{WP}$ is the usual Weil-Petersson metric on ${\mathfrak M}_g$. This is essentially (\ref{8.5}). Furthermore, Wolpert \cite{Wo} has shown that $\omega_{WP}$ extends to a closed form with singularities on the campactification $\overline{\mathfrak M}_g$ of the moduli space and in the sense of currents, $\frac{1}{\pi^2}\omega_{WP}$ is the Chern form of a continuous metric $h_{WP}$ on a certain line bundle $\lambda_{WP}$. This gives a projective embedding of $\overline{\mathfrak M}_g$.
\section{Introduction} \label{sec:intro} One of the major problems in the theory correlated electrons is to construct in a systematic and controlled way a consistent approximation interpolating reliably between weak- and strong-coupling regimes. The two extreme limits of weak and strong couplings in the archetypal Hubbard model can be described relatively well. The weak-coupling regime is governed by a Hartree-Fock mean field with dynamical fluctuations covered by Fermi-liquid theory. Extended systems at low temperatures are Pauli paramagnets with smeared out local magnetic moments. For bipartite lattices antiferromagnetic long-range order sets in at half filling and zero temperature at arbitrarily small interaction. In the strong-coupling regime the Hubbard model at half filling maps onto a Heisenberg antiferromagnet with pronounced local magnetic moments and the Curie-Weiss law for the staggered magnetic susceptibility, at least at the mean-field level. The spectral structure is dominated by separated lower and upper Hubbard bands and the strongly correlated system seems to be insulating even in the paramagnetic phase. However, it is the intermediate coupling, where the effective Coulomb repulsion is comparable with the kinetic energy and hence neither very weak nor very strong, that is of great interest for the theorists as well as for the experimentalists. At intermediate coupling, dynamical fluctuations control the low-temperature physics of interacting electrons and neither weak-coupling nor atomic-like perturbation theories are adequate. In this nonperturbative regime a singularity in a two-particle function is approached and we expect breakdown of the Fermi-liquid regime and a transition to an ordered state breaking the symmetry of the Hamiltonian. Most of the theoretical effort in strongly correlated electron systems concentrates on finding an appropriate and reliable description of this fluctuation-dominated transition region. There are only a few general theoretical means with which we could address the weak-to-strong-coupling transition. The only exact, Bethe-ansatz solution in $d=1$ \cite{Lieb68} does not have a weak-coupling regime and numerical techniques such as quantum Monte-Carlo simulations or exact diagonalization are restricted to relatively high temperatures or small clusters \cite{Dagotto94}. They cannot cope efficiently with the very different energy scales relevant in the transition regime. Recent progress in the dynamical mean-field theory constructed from the limit $d\to\infty$, where only frequency fluctuations are dynamically relevant, has brought new insights into possible scenarios of the transition from weak to strong couplings. The transition at the mean-field level is connected with a Kondo-like behavior and the Mott-Hubbard metal-insulator transition. Its critical behavior has not yet been fully cleared. A widely accepted picture for the transition at zero temperature, initially based on the so-called iterated perturbation theory (IPT) \cite{Zhang93} and lately supported by analytical \cite{Moeller95} and numerical \cite{Bulla98} renormalization-group arguments, was, however, recently questioned using assessments from perturbation theory and a skeleton expansion \cite{Kehrein98}. Unfortunately even in this particular limit of high spatial dimensions, no exact solution to the Hubbard model has yet been found and our predictions and conclusions are justified by approximate schemes of either numerical or analytical character \cite{Georges96}. The need for a nonperturbative quantitative scheme being stable throughout the whole range of the interaction strength is still extreme. We concentrate in this paper exclusively on analytically controllable approximations based on partial summations and renormalizations of Feynman diagrams. Only these approaches have the power to give the details of the critical behavior of correlated electrons. To decide which of the existing approaches are prospective in a reliable description of quantum critical behavior, we investigate in detail stability of approximations constructed within renormalized weak-coupling expansions continued to intermediate and strong couplings. We use the framework of Baym and Kadanoff \cite{Baym61,Baym62} to formulate general stability criteria that must be imposed upon any approximate theory intending to be thermodynamically consistent and conserving. I.~e. we use an expansion with renormalized one-electron propagators and construct a Luttinger-Ward generating functional from which all the physical quantities are derived via functional derivatives. Additively we use small external perturbations to test stability of the equilibrium states. In this paper we extend the Baym and Kadanoff scheme in two important aspects. First, we generalize external perturbations of the equilibrium system from density fluctuations also to anomalous sources not conserving charge and/or spin. Only in this way we control all relevant two-particle vertex functions that may show divergences and hence lead to instabilities. We use the disturbances as generators of two-particle irreducibility and find a complete set of criteria for a local stability of self-consistent solutions with a rigorous relation between singularities in two-particle functions, indicating an instability of the solution, and symmetry-breaking order parameters. Second, since two-particle functions play extremely important role in the stability analysis we have to treat the vertex functions and the one-electron propagators on the same footing within a single renormalization scheme. We must not loose any important relation between various two-particle quantities as well as between one- and two-particle functions. This is achieved if instead of approximating the self-energy functional we shift diagrammatic approximations to the level of two-particle functions. Equations of motion for one- and two-particle Green functions are essential in our construction of approximate theories. They are the Schwinger-Dyson and Bethe-Salpeter equations. The former connects the one-particle self-energy with the full two-particle vertex function. The latter equations are defined separately for each two-particle irreducibility channel and determine finally the full vertex as a functional of the completely two-particle irreducible one. Since we expect that the Bethe-Salpeter equations may get singular and the full vertex function can have poles at intermediate and strong coupling, we do not attempt to destroy the structure of these equations of motion. Instead we introduce diagrammatic approximations only in the input to these equations, i.~e. in the completely irreducible two-particle vertex. Thereby we do not miss possible singularities of the vertex function assumed the relevant physics is essentially determined by two-particle scatterings. Such a renormalization of Feynman diagrams, or construction of skeleton diagrams, called parquet approach \cite{Dominicis62,Jackson82}, is from the very beginning self-consistent at the level of one- as well as two-particle functions. It contains dynamical vertex renormalizations and guarantees that divergences in two-particle functions are treated properly. The parquet approach or the parquet diagrams are known for long in many-body physics. First introduced in nuclear physics \cite{Sudakov56}, the parquet theory found quickly its way into condensed matter. It was attempted to solve the Kondo problem \cite{Abrikosov64}, the x-ray edge \cite{Roulet69}, or the formation of the local magnetic moment in dilute alloys \cite{Weiner70} with the parquet approximation. However, in none of these applications a full solution to the parquet equations was found. In the former two approaches an expansion with leading-logarithmic divergences of single electron-hole bubbles was used. This construction underestimates the role of the genuine poles in the Bethe-Salpeter equations due to multiple two-particle scatterings. It actually amounts to a loop expansion neglecting the complex dynamics of the vertex functions. The last approach resorted to static approximations and had only a restricted success in the strong-coupling limit. It is the intricate structure of the parquet equations that makes nonperturbative solutions inavailable and hinders a broader application of the parquet construction. Having a general scheme for deriving theories with renormalized one- and two-particle functions we must reach for approximations to come to quantitative results. The simplest theory fulfilling the equations of motion for both one- and two-particle Green functions is the parquet approximation where the completely irreducible vertex is just the bare interaction. However, up to now no nonperturbative solution to the parquet equations has yet been found. There already exist simplifications of the parquet construction (vertex renormalization) \cite{Bickers91,Vilk94}, but because they resort to static vertices they get unstable and fail at strong coupling. Here we reduce the intricate complex structure of the full parquet approximation so that the vertex functions are dynamically renormalized and allow only for integrable divergences. We demand an asymptotic accuracy of the simplification at critical points and keep only maximally divergent contributions to the vertex function \cite{Janis98a,Janis99a}. We end up at zero temperature and half filling with a theory stable throughout the whole range of the interaction strength the computational complexity of which is comparable with the single-channel or FLEX-type approximations. We apply it to a mean-field description of the Mott-Hubbard metal insulator transition. The layout of the paper is as follows. Section~\ref{sec:stability} presents the general scheme for deriving approximations with renormalized one- and two-particle propagators allowing us to formulate stability criteria with the relevant two-particle functions. We explicitly show in Sections~\ref{sec:hartree},~\ref{sec:second-order} that all approximations simpler than the parquet diagrams get unstable at intermediate coupling of the Hubbard and single-impurity Anderson models. Stability criteria for the parquet approximation are explicitly formulated in Sec.~\ref{sec:parquet}. The proposed simplification of the parquet approximation with dynamical vertex corrections is presented in Sec.~\ref{sec:simplifications}. We demonstrate stability of the simplified parquet approach in the mean-field theory of the Mott-Hubbard metal-insulator transition in Sec.~\ref{sec:DMFT}. Last Section~\ref{sec:conclusions} brings discussion and concluding remarks. \section{Equations of motion, two-particle Green functions and stability of solutions} \label{sec:stability} We use two generic Hubbard and single-impurity Anderson Hamiltonians with an external magnetic field $B$ to model moderate and strong electron correlations. The magnetic field allows us to lift the degeneracy in the spin space and to determine the relevant two-particle functions we need to look at if we want to prove stability of the theory. We start with equilibrium Hamiltonians \begin{mathletters} \begin{eqnarray} \label{eq:hh} \widehat{H}_H&=&\sum_{{\bf k}\sigma} \left(\epsilon({\bf k}) +\mu+\sigma B\right) c^{\dagger}_{{\bf k}\sigma} c^{\phantom{\dagger}}_{{\bf k}\sigma} + U\sum_{{\bf i}}\widehat{n}_{{\bf i}\uparrow}\widehat{n}_{{\bf i} \downarrow}, \\ \label{eq:ah} \widehat{H}_A&=&\sum_{{\bf k}\sigma} \left(\epsilon({\bf k}) +\sigma B\right) c^{\dagger}_{{\bf k}\sigma}c^{\phantom{\dagger}}_{{\bf k}\sigma} +\sum_{{\bf k}\sigma} \left(V_{{\bf k}} c^{\dagger}_{{\bf k}\sigma}f_\sigma +V_{{\bf k}}^* f^\dagger_\sigma c^{\phantom{\dagger}}_{{\bf k}\sigma}\right) +\sum_\sigma(\varepsilon_f +\sigma B) f^\dagger_\sigma f_\sigma +U\widehat{n}_{f\uparrow}\widehat{n}_{f\downarrow} . \end{eqnarray} \end{mathletters} The Hubbard Hamiltonian (\ref{eq:hh}) is of primary interest for us, since phase transitions with long-range order can appear there. The Anderson Hamiltonian (\ref{eq:ah}) is considered only to demonstrate the difference between a mean-field solution of the Hubbard model ($d=\infty$) and the impurity solution and to test adequacy of chosen approximate schemes. The most direct way to study stability of solutions at thermal equilibrium is to introduce auxiliary sources as an external perturbation \cite{Baym61}. Only external sources conserving spin and charge leading to density-density (nonanomalous) response functions are standardly used to perturb the equilibrium state. However, these sources are not complete and do not allow for anomalous responses and quantum phases such as superconductivity. To be able to describe all possible instabilities of the Hubbard-like models we introduce a generalized external perturbation \begin{eqnarray} \label{eq:H-ext} \widehat{H}_{ext}&=&\int d1d2\Bigg\{\sum_\sigma\left[\eta^{||}_\sigma(1,2) c^{\dagger} _{\sigma}(1) c^{\phantom{\dagger}}_{\sigma}(2)+\bar{\xi}^{||}_\sigma (1,2)c^{\phantom{\dagger}}_{\sigma}(1)c^{\phantom{\dagger}}_{\sigma} (2)+\xi^{||}_\sigma(1,2) c^{\dagger}_{\sigma}(1)c^{\dagger} _{\sigma}(2)\right]\nonumber\\ &&+\left[\eta^{\perp}(1,2)c^{\dagger}_{\uparrow}(1)c^{\phantom{\dagger}} _{\downarrow}(2) +\bar{\eta}^{\perp}(1,2)c^{\dagger}_{\downarrow}(2) c^{\phantom{\dagger}}_{\uparrow}(1)\right]+\left[\bar{\xi}^{\perp} (1,2)c^{\phantom{\dagger}}_{\uparrow}(1)c^{\phantom{\dagger}} _\downarrow(2)+\xi^{\perp}(1,2)c^{\dagger}_{\downarrow}(2) c^{\dagger}_{\uparrow}(1)\right]\Bigg\} \end{eqnarray} where labels $1=({\bf r}_1,t_1)$, $2=({\bf r}_2,t_2)$ denote a set of space-time variables characterizing the motion of quantum particles, electrons. Here the local real fields $\eta^{||}_\sigma$ induce density responses conserving charge as well as spin. These fields (e.~g. magnetic) are used in standard stability analyses. The new nonconserving external sources are complex and either add or remove charge, spin or both from the equilibrium many-particle state. The field $\eta^{\perp}$ conserves charge but increases spin of the equilibrium state by two elementary units. The fields $\xi^{\perp}$ increase charge while $\xi^{||}_\sigma$ increase both charge and the appropriate spin projection. The complex conjugate fields lower the respective quantities. Fundamental quantity for quantum statistics is a thermodynamic potential as a functional of the external perturbation $\widehat{H}_{ext}$. We need to expand the generating thermodynamic potential only to second order in the external perturbation in order to determine (local) stability of the equilibrium state. First term of this expansion generates one-particle Green functions. The real fields lead to regular propagators, while the complex ones to anomalous Green functions. They vanish at equilibrium unless we are in a quantum phase with anomalous Green functions. To decide whether a long-range order may arise or whether the equilibrium state is stable we have to evaluate second term in the external perturbation. It defines two-particle functions, regular and anomalous. In the unperturbed, equilibrium state without anomalous functions only diagonal terms in the external sources do not vanish. If $H_\alpha$ denotes a chosen external source perturbing the equilibrium system, i.~e. it stands generically for $\eta^{||}$, $\eta^{\perp}$, $\xi^{||}$, and $\xi^{\perp}$, we can write \begin{eqnarray} \label{eq:G2} G^{(2)\alpha}(13,24)&=&\frac{\delta^2\Phi[G,H]}{\delta H_{\alpha}(4,3) \delta H_{\bar{\alpha}}(2,1)}\Bigg|_{H=0}\ . \end{eqnarray} It is interesting to note that second derivatives w.r.t. different external sources introduced in perturbation $\widehat{H}_{ext}$, (\ref{eq:H-ext}), lead at equilibrium to inequivalent two-particle irreducibility channels, diagrams that cannot be made disconnected by cutting two specific lines. More precisely, the external sources generate two-particle reducible functions, while their Legendre conjugate ``order parameters'' the respective irreducible functions. The field $\eta^{||}$ leads to reducible functions in the interaction ($U$) channel and evokes normal density responses and susceptibilities. The field $\eta^{\perp}$ leads to the electron-hole ($eh$) and $\xi^{\perp}$ to the electron-electron ($ee$) reducible functions, respectively. Second derivative w.r.t. $\xi^{||}$ vanishes at equilibrium without anomalous functions \cite{note0}. It is more convenient to introduce a vertex function instead of the full two-particle function $G^{(2)\alpha}$. We subtract the free two-particle propagation, if contributes, and stripe off the external legs from the remaining two-particle function. The result is a vertex function $\Gamma$ independent of the index $\alpha$ \cite{note1}. The full vertex function $\Gamma$ can be decomposed equivalently in each two-particle channel into a sum of the irreducible and reducible functions $\Lambda$ and $\mathcal{K}$, respectively: \begin{eqnarray} \label{eq:vertex-def} \Gamma&=&\Lambda^\alpha+\mathcal{K}^\alpha . \end{eqnarray} To go on we have to introduce a convention for independent variables labeling the two-particle functions. We denote $X_{\sigma\sigma'}(k,k';q)$ the generic (regular) two-particle function in the momentum representation. We use four-momenta, $k=({\bf k},i\omega_n)$ for the fermionic and $q=({\bf q},i\nu_m)$ for the bosonic variables with Matsubara frequencies $\omega_n=(2n+1)\pi$, $\nu_m=2m\pi$. The meaning of each variable in the function $X$ is manifested in Fig.~\ref{fig:2P-generic}. The incoming arrows denote annihilation of charge. \begin{figure} \hspace*{80pt} \epsfig{figure=parqfig0.eps,height=40mm} \caption{\label{fig:2P-generic} Generic two-particle function with three independent four-momenta and a defined order of incoming and outgoing fermions.} \end{figure} A two-particle irreducible function in a specified channel $\alpha$ can be defined as a functional derivative of the self-energy $\Sigma^{\alpha}$ from the $\alpha$-channel with respect to the renormalized one-electron propagator \begin{mathletters}\label{eq:IGF-diff} \begin{eqnarray} \label{eq:2IP-diff} \Lambda^\alpha(13,24)&=&\frac{\delta\Sigma^\alpha(1,2)}{\delta G^\alpha(4,3)} \Bigg|_{H=0} \ , \end{eqnarray} where the self-energy is a functional of the full one-electron propagator defined from the generating Luttinger-Ward functional \begin{eqnarray} \label{eq:1IP-diff} \Sigma^\alpha(1,2)&=&\frac{\delta\Phi[G,H]}{\delta G^\alpha(2,1)}\ . \end{eqnarray}\end{mathletters} Note that even if the anomalous self-energy $\Sigma^{\alpha}$ vanishes at equilibrium the two-particle function is nontrivial, since it is conserves both charge and spin. Up to now we looked at the equilibrium system from the thermodynamic point of view. We assumed we have a Luttinger-Ward generating functional $\Phi[G]$ from which we determine all the relevant quantities via functional derivatives. However, we do not know how to find the Luttinger-Ward functional. A naive way is to choose a particular set of skeleton diagrams directly for the grand potential. A more convenient and usual way is to choose diagrams at the one-particle level by fixing the functional $\Sigma[G]$. However, in neither of these constructions we have a connection to the fundamental equations of motion for the one- and two-particle Green functions. The equations of motion determine the microscopic dynamics of quantum systems. If we want to simulate the exact dynamics of the studied system we must try to approximate the equations of motion as accurate as possible. First equation of motion we have to fulfill is the Schwinger-Dyson equation connecting the one-electron self-energy with the vertex function. It reads for Hubbard-like models with a local electron-electron interaction: \begin{eqnarray} \label{eq:sigma-2P} \Sigma_\sigma(k)&=&\frac{U}{\beta N}\sum_{k'} G_{-\sigma}(k') -\frac{U} {\beta^2N^2}\sum_{k'q}\Gamma_{\sigma-\sigma}(k,k';q) G_\sigma(k+q)G_{-\sigma}(k'+q)G_{-\sigma}(k'). \end{eqnarray} The first term on the right-hand side is the Hartree self-energy and the latter contains second and higher-order, dynamical corrections. The usual way to deal with the Schwinger-Dyson equation is to introduce approximations to the vertex function $\Gamma[G]$ on the right-hand side of (\ref{eq:sigma-2P}). We thereby get a closed functional $\Sigma[G]$ and close the approximation \cite{Baym62}. However, at the critical points, where equilibrium states get unstable, the full vertex function $\Gamma$ is singular. Brute interferences in the Schwinger-Dyson equation can make the approximate dynamics inadequate and qualitatively different from the exact one. Since we are primarily interested in the two-particle functions and their appropriate description, we move approximations from the full vertex function to a lower level, namely to the completely irreducible two-particle function. Two-particle irreducible vertices determine the full vertex function via Bethe-Salpeter equations. These equations explicitly sum two-particle reducible diagrams in each two-particle channel. In the notation of Fig.~\ref{fig:2P-generic} we obtain \begin{mathletters}\label{eq:Bethe-Salpeter} \begin{eqnarray} \label{eq:2P-reducible} \Gamma(k,k';q)&=&\Lambda^\alpha(k,k';q)-\left[\Lambda^\alpha GG\odot \Gamma \right](k,k';q) . \end{eqnarray} The Bethe-Salpeter equations must be completed by definitions connecting the respective irreducible and reducible functions $\Lambda^\alpha$, ${\cal K}^\alpha$, respectively. Due to inequivalence of different two-particle channels we have \begin{eqnarray} \label{eq:2IP-def} \Lambda^\alpha&=& I+\sum_{\alpha'\neq\alpha}{\cal K}^{\alpha'} \end{eqnarray} \end{mathletters} where $I[U;G,\Gamma]$ is the completely irreducible two-particle vertex being a functional of the one-particle propagator $G$ and generally also of the vertex function $\Gamma$. When expressed diagrammatically, the completely irreducible vertex can be disconnected only by cutting at least three one-electron propagators. We used a generic notation $\odot$ for the channel-dependent multiplication of two-particle functions. It mixes the variables of two-particle functions in different manners. The three matrix multiplication schemes for two-particle quantities in our notation of Fig.~\ref{fig:2P-generic}, representing summations over intermediate states, read in the interaction, electron-hole, and electron-electron channels, respectively \begin{mathletters}\label{eq:conv} \begin{eqnarray} \label{eq:conv-U} \left[\widehat{X}GG\star\widehat{Y}\right]_{\sigma\sigma'}(k,k';q)&=&\frac 1{\beta{\cal N}}\sum_{\sigma''k''}X_{\sigma\sigma''}(k,k'';q) G_{\sigma''} (k'')G_{\sigma''}(k''+q)Y_{\sigma'' \sigma'}(k'',k';q), \\ \label{eq:conv-eh} \left[\widehat{X}GG\bullet\widehat{Y}\right]_{\sigma\sigma'}(k,k';q)&=& \frac 1{\beta{\cal N}}\sum_{q''} X_{\sigma\sigma'}(k,k';q'')G_{\sigma} (k+q'')G_{\sigma'}(k'+q'')\nonumber \\&& \hspace*{40pt}\times Y_{\sigma\sigma'}(k+q'',k'+q'';q-q''),\\ \label{eq:conv-ee} \left[\widehat{X}GG\circ\widehat{Y}\right]_{\sigma\sigma'}(k,k';q)&=&\frac 1{\beta{\cal N}}\sum_{q''} X_{\sigma\sigma'}(k,k'+q'';q-q'')G_{\sigma} (k+q-q'') G_{\sigma'}(k'+q'')\nonumber\\ && \hspace*{40pt}\times Y_{\sigma\sigma'}(k+q-q'',k';q'') . \end{eqnarray} \end{mathletters} For convenience we used separate symbols for each multiplication scheme. Note that only the interaction (vertical) channel mixes the spin singlet and triplet functions. Now all the quantities depend explicitly only on the model input parameters and are functionals of the completely irreducible vertex $I[U;G,\Gamma]$ to which approximations can be chosen. Two-particle functions are determined self-consistently from (\ref{eq:Bethe-Salpeter}) and (\ref{eq:conv}). The one-electron self-energy is defined from (\ref{eq:sigma-2P}). The generating Luttinger-Ward functional is determined from the functional differential equation (\ref{eq:1IP-diff}). Its explicit form for the parquet approximation with $I=U$ is derived in the Appendix. Approximate theories break some exact relations and the thermodynamic and dynamical definitions of the two-particle functions do not coincide. If we enter an approximation, we have ambiguous definitions for two-particle irreducible functions. They can be defined either from (\ref{eq:IGF-diff}) or from (\ref{eq:Bethe-Salpeter}). This is an unrecoverable feature of all approximate theories which one has to live with. The only thing one can demand from reliable approximations is that both the definitions of two-particle functions lead to qualitatively the same physical behavior (phase diagram and spectral properties). The vertex functions from the equations of motion are responsible for spectral and analytic properties of the approximate theory and are used to determine internal stability (consistency) of approximations. On the other hand the differential definition of the vertex functions is used for determining thermodynamic behavior and stability with respect to external disturbances. They may be either due to external sources or due to neglected dynamical vertex renormalizations. The advantage of the above general formulation with equations of motion is that we treat the one- and two-particle functions consistently on the same footing within one renormalization scheme. This is essential for determining stability of approximations. We now rewrite the channel-dependent matrix multiplications so as to distinguish manifestly active and conserved variables (parameters). We introduce new notations in which only the active four-momenta are used as variables. We define the following symbols \begin{mathletters} \label{eq:channels} \begin{eqnarray} \label{eq:U} \left[X GG\right]^U[q](\sigma k,\sigma'k')&=&X_{\sigma \sigma'}(k,k';q)G_{\sigma'}(k')G_{\sigma'}(k'+q),\\ \label{eq:eh} \left[X GG\right]^{eh}_{\sigma\sigma'}[q] (k,k')&=&X _{\sigma\sigma'}(k,k+q,k'-k) G_{\sigma}(k')G_{\sigma'}(k'+q),\\ \label{eq:ee} \left[X GG\right]^{ee}_{\sigma\sigma'}[q](k,k')&=&X _{\sigma\sigma'}(k,k';q-k-k')G_{\sigma}(k')G_{\sigma'}(q-k') . \end{eqnarray} \end{mathletters} In (\ref{eq:channels}) the bosonic variable $q$ is always inactive in the multiplication, or conserved during the scatterings within the chosen two-particle channel. Each channel, however, has a different conserving variable. It makes the parquet algebra of two-particle functions complicated. We can, in principle, find eigenvalues and eigenvectors for the matrices, kernels of the Bethe-Salpeter equations (\ref{eq:Bethe-Salpeter}), $[\Lambda^\alpha GG]^\alpha$, and denote them $Q^\alpha$. These are complex four-vectors. Only real four-vectors, i.e. for zero frequency, both the conserved $q^\alpha$ as well as the eigenvectors $Q^\alpha$ are important for determining the stability conditions. They may formally be written as \begin{mathletters} \label{eq:stability} \begin{eqnarray} \label{eq:stability1} \mbox{min}\left[\Lambda^{\alpha}GG\right]^\alpha[\textbf{q}^\alpha,0] (\textbf{Q}^\alpha,0) &\ge& -1 \end{eqnarray} or equivalently using (\ref{eq:2IP-def}) \begin{eqnarray} \label{eq:stability2} \mbox{min}\left[\left(I+\sum_{\alpha'\neq\alpha}{\cal K}^{\alpha'} \right)GG\right]^\alpha[\textbf{q}^\alpha,0](\textbf{Q}^\alpha,0)&\ge&-1 . \end{eqnarray}\end{mathletters} This is a complete set of conditions for a local stability of any symmetric solution fulfilling the Bethe-Salpeter equations of motion. The inequalities are a direct consequence of the nonexistence of singularities at the real axis in equations (\ref{eq:Bethe-Salpeter}). Equality in (\ref{eq:stability}) indicates that a pole in the respective two-particle (reducible) Green function appears and consequently a singularity in a correlation function such as susceptibility. Due to (\ref{eq:2IP-def}), the singularity is transferred to the irreducible functions from the other two-particle channels and unless the singularity is {\em integrable}, the scatterings with singular irreducible kernels destroy the transition. We hence can have only integrable singularities in the exact theory. A singularity in an approximate two-particle function means on one hand the existence of long-range correlations as a precursor of a long-range order, but on the other hand also signals critical vertex corrections stabilizing the symmetric phase. These two critical phenomena compete and we must compare the symmetry breaking contributions at the critical point with the singular vertex renormalization in approximate theories \cite{Janis98a}. The nonexistence of poles at the real axis is a condition for the thermodynamic stability of the symmetric (paramagnetic) phase. However, a consistent solution for $\Sigma$ and $\Gamma$ must produce analytic functions in complex frequencies. This is achieved if the analytic continuations of the two-particle functions from Matsubara frequencies have no poles in the complex plane away from the real axis. Hence the proper analyticity of the solution is guaranteed if \begin{eqnarray} \label{eq:stabibilty3} \left[\Lambda^{\alpha}GG\right]^\alpha[{\bf q}^\alpha,z]({\bf Q}^\alpha, Z)&\neq& -1 \end{eqnarray} where $\mbox{Im} z$,$\mbox{Im} Z \neq 0$. If these conditions are fulfilled the self-energy determined from (\ref{eq:sigma-2P}) is analytic in the lower and upper complex half-planes. The stability criteria, when applied to approximate solutions, must be used with both two-particle functions. The two-particle functions from the equations of motion are used to determine internal consistency of the approximation and especially (\ref{eq:stabibilty3}) is used to prove analyticity of the approximate Green functions. The thermodynamically derived vertex functions are used to check stability of the approximation with respect to fluctuations caused either by long-range pair correlations or by neglected vertex corrections. An ``external'' instability of an approximation hence need not automatically mean a transition to an ordered state. It may happen for oversimplified approximations, what actually is the case in low dimensions, that the neglected dynamical vertex renormalization smears out the transition. Hence, before we make conclusions on the existence of a symmetry breaking with a long-range order, we must be sure the neglected vertex corrections are irrelevant. If not, the approximate LRO becomes unreliable and further vertex renormalizations are indispensable. \section{Hartree approximation} \label{sec:hartree} We investigate in this and the following sections to what degree various approximations produce stable equilibrium states at intermediate and strong coupling. Only stable solutions are reliable or suitable for the description of the transition from weak to strong couplings. We start with the simplest, Hartree approximation. The Hartree approximation is obtained from our construction if we put $\Gamma=0$ in (\ref{eq:sigma-2P}). The theory is internally consistent in a trivial way and the only nontrivial two-particle function is that obtained from (\ref{eq:IGF-diff}). To determine the external stability we have to evaluate single two-particle bubbles \begin{eqnarray} \label{eq:2P-bubble} X_{\sigma\sigma'}({\bf q},i\nu_m)&=&\frac U{\beta{\cal N}}\sum_{{\bf k}n} G_\sigma({\bf k},i\omega_n) G_{\sigma'}({\bf k+q},i(\omega_n+\nu_m)) \end{eqnarray} with the Hartree propagators. They explicitly read for the Hubbard and single-impurity models \begin{mathletters}\label{eq:GF-def} \begin{eqnarray} \label{eq:GF-Hubbard} G_\sigma({\bf k},z)&=&\left[z+\mu+\sigma B-Un_{-\sigma}-\epsilon({\bf k})\right]^{-1} , \\ \label{eq:GF-SIAM} G_\sigma(z)&=&\left[z-\varepsilon_f+\sigma B-Un_{-\sigma}-V^2g(z+\sigma B)\right]^{-1} \end{eqnarray} \end{mathletters} where $g(z)$ is the local element of the one-electron propagator of the conduction electrons. We use $g(z)=2(z-\sqrt{z^2-1})/w^2$ corresponding to a Bethe lattice in $d=\infty$. We evaluate the bubbles for the Hubbard and single impurity models only in the paramagnetic phase at half filling, the electron-hole symmetric situation. We start with arbitrary finite dimension. The singlet and triplet electron-hole bubbles after summation over Matsubara frequencies are: \begin{mathletters} \label{magnetic.bubbles} \begin{eqnarray} X_{\uparrow\downarrow}({\bf q},\zeta)&=&X_{\downarrow\uparrow} (-{\bf q},-\zeta) =\frac U{\cal N}\sum_{\bf k}\frac{f\left(\epsilon({\bf k}) -B-Um/2\right)-f\left(\epsilon({\bf k}+{\bf q}) +B+ Um/2\right)}{\zeta -2B-Um +\epsilon({\bf k})-\epsilon({\bf k}+{\bf q})}, \\ X_{\sigma\sigma}({\bf q},\zeta)&=& \frac U{\cal N}\sum_{\bf k}\frac{f\left(\epsilon({\bf k}) -\sigma(B+ Um/2)\right)-f\left(\epsilon({\bf k}+{\bf q})-\sigma(B+Um/2) \right)} {\zeta+\epsilon({\bf k})-\epsilon({\bf k}+{\bf q})} . \end{eqnarray} \end{mathletters} The Hartree solution becomes unstable if \begin{eqnarray} \label{eq:HF-instab} 1+X_{\sigma\sigma'}({\bf q},0)&=&0 . \end{eqnarray} This happens first for the singlet $X_{\uparrow\downarrow}$ bubble at a critical field $B_{c0}$. It can be understood as a separator between the weak and intermediate coupling regimes, since iterations from the Hartree solution get unstable beyond the stability limit, Fig.~\ref{fig:hartree-d3}. \begin{figure} \epsfig{figure=hartree_d3.ps,height=12cm, angle=-90} \caption{\label{fig:hartree-d3} Instability of the Hartree approximation for the half-filled Hubbard model in an external magnetic field for low temperatures in $d=3$. The boundary field $B_{c0}$ separates weak and intermediate couplings (WC, IC). The energy unit is fixed by the half-bandwidth $w=1$.} \end{figure} The critical field $B_{c0}$ coincides at zero temperature with the boundary to the saturated ferromagnetic solution $B_{s}$ \cite{vanDongen94}. This holds even in the weak-coupling regime with no singularity in the fully saturated ferromagnetic state. For if we get below the line of instability of the fully saturated ferromagnet, defined by $B_s=w-U/2$, the left-hand side of (\ref{eq:HF-instab}) becomes negative and logarithmically diverges with a nonvanishing weight $\rho(B+Um/2)$, where $\rho$ is the density of states and $m<1$. We now perform the same analysis for the SIAM in a Bethe lattice and the Hubbard model in $d=\infty$ dimensions where only frequency variables are relevant and no long-range order can arise. However, the strong-coupling phase may get insulating. We choose the following value of the hybridization $V=w/\sqrt{2}$ with the energy unit set by the energy half-bandwidth $w=1$. The electron-hole bubble of an impurity embedded in a lattice with a finite bandwidth has two contributions, band and impurity ones. They are at zero temperature \begin{mathletters} \begin{eqnarray} \label{eq:HF-SIAM-gamma-b} \Gamma(0)^b_{\uparrow\downarrow}&=&-U\int_{-1}^{-1+2B}\frac{dx}{\pi}\ \frac{\sqrt{1-x^2}}{U^2m^2/4+1-x^2}\ \frac 1{Um/2+\sqrt{(x-2B)^2-1}} \nonumber\\ &&-\frac{U^2m}2\int_{-1+2B}^B\frac{dx}{\pi}\ \frac{\sqrt{1-x^2}+\sqrt{1- (x-2B)^2}}{\left[U^2m^2/4+1-x^2\right]\left[U^2m^2/4+1-(x-2B)^2 \right]} , \end{eqnarray} \begin{eqnarray} \label{eq:HF-SIAM-gamma-I} \Gamma(0)^I_{\uparrow\downarrow}&=&-\frac{U^2m}{2\sqrt{1+U^2m^2/4}}\ \frac 1{Um/2+\sqrt{\left(2B+\sqrt{1+U^2m^2/4}\right)^2-1}} \end{eqnarray} \end{mathletters} where the local magnetization is \begin{eqnarray} \label{eq:HF-SIAM-m} m&=&\frac{Um}{2\sqrt{1+U^2m^2/4}}+\int_{-B}^B\frac{dx}\pi \frac{\sqrt{1-x^2}}{U^2m^2/4+1-x^2} . \end{eqnarray} Evaluating the instability condition $\Gamma^b(0)+\Gamma^I(0)=-1$ we obtain a critical field beyond which the Hartree approximation is no longer stable. It is plotted at zero temperature in Fig.~\ref{fig:hartree-imp}. The boundary field at which the fully polarized solution ($m=1$) breaks down is independent of the interaction strength, $B_s=w$, \cite{Janis99b}. The Hartree solution becomes exact in the fully spin-polarized state. At $B=0$, the instability of the Hartree solution with respect to a spin flip, $U_{c0}$, lies below the interaction $U_F=2w$ at which the solution breaks into a spin polarized state with $m\neq0$. \begin{figure} \epsfig{figure=hartree_imp.ps,height=12cm, angle=-90} \caption{\label{fig:hartree-imp} Instability of the Hartree approximation for the half-filled SIAM in an external magnetic field at zero temperature. The dashed line is the local magnetization along the instability line $m_{c0}$.} \end{figure} A mean-field solution with only frequencies (i.e. neglecting any long-range order) behaves differently. The singlet electron-hole bubble for a Bethe lattice reads \begin{mathletters} \begin{eqnarray} \label{eq:HF-MF-gamma} \Gamma(0)_{\uparrow\downarrow}&=&-\frac{8U}\pi\left\{\frac 23\left[1 -\left(B+\frac U2m\right)^2\right]^{3/2}+\frac \pi2\left(B+\frac U2m\right)m\right\}\nonumber\\[2pt] && +4U\int_{-1}^{-1+2B+Um}\frac{dx}\pi\sqrt{1-x^2}\sqrt{(2B+Um-x)^2-1} \end{eqnarray} with the magnetization defined by an equation \begin{eqnarray} \label{eq:HF-MF-m} m&=&\frac 2\pi\left[\arcsin\left(B+\frac U2 m\right)+\left(B+\frac U2 m \right)\sqrt{1-\left(B+\frac U2 m\right)^2}\ \right] . \end{eqnarray} \end{mathletters} The boundary field for the fully polarized solution is now a function of the interaction strength alike in finite-dimensional models, Fig.~\ref{fig:bcrit-mf}. The instability of the Hartree approximation outside the fully polarized solution is plotted in Fig.~\ref{fig:hartree-mf}. Unlike finite-dimensional cases, the weak-coupling regime is robust at zero temperature, since we suppressed any possible long-range order. At $B=0$, the instability of the Hartree solution appears at $U_{c0}=3\pi w/8<U_F=\pi w/2$. It lies deep below the expected metal-insulator transition at $U_c\approx 3w$ \cite{Georges96}. It means that the Hartree approximation for effective impurity models gets unstable and unreliable before a spin-polarized or insulating solution may appear. We must go to more sophisticated approximations if $U>w$ and we want to approach the expected metal-insulator transition. \begin{figure} \epsfig{figure=bcrit_mf.ps,height=12cm, angle=-90} \caption{\label{fig:bcrit-mf} Boundary field $B_s$ for the fully polarized solution of the Hubbard model at $T=0$ on a $d=\infty$ Bethe lattice with local two-particle functions. The boundary field for single impurity is $B_s=w=1$.} \end{figure} \begin{figure} \epsfig{figure=hartree_mf.ps,height=12cm, angle=-90} \caption{\label{fig:hartree-mf} Instability of the Hartree approximation for the half-filled mean-field solution of the Hubbard model in an external magnetic field at zero temperature. The upper region (SF) is a fully polarized state ($m=1$).} \end{figure} \section{Second order and single-channel approximations} \label{sec:second-order} An instability in the Hartree approximation in the electron-hole channel, i.e. for the electron-hole singlet bubble, indicates either the existence of a symmetry breaking with anomalous functions (in an external magnetic field) or an instability with respect to dynamical fluctuations. They can smear out the static Hartree instability and keep the symmetric solution stable. We now check stability of simple extensions of the Hartree mean-field theory. \subsection{Self-consistent second order} \label{sec:SCSOPT} First step including dynamical fluctuations beyond Hartree is second-order perturbation theory. We obtain this approximation if we replace the full vertex function $\Gamma$ in (\ref{eq:sigma-2P}) by the bare coupling constant $U$. Since the right-hand side of (\ref{eq:sigma-2P}) contains no infinite series, the only stability criterion for this approximation is that with (\ref{eq:2IP-diff}). Alike the Hartree approximation we expect that the instability arises first in the electron-hole channel with spin singlet bubbles \begin{eqnarray} \label{eq:SOPT-vertex} \left[\frac{\delta\Sigma}{\delta G}GG\right]^{eh}_{\sigma-\sigma}[q](k,k') &=& U(1-X_{\sigma-\sigma}(q))G_{\sigma}(k')G_{-\sigma}(k'+q) \end{eqnarray} where $X_{\sigma\sigma'}(q)$ was defined in (\ref{eq:2P-bubble}). We have to find the lowest real eigenvalue of the above expression seen as a matrix in $k$ and $k'$. Since the matrix does not explicitly depend on $k$, the lowest eigenvalue is just a sum over the other variable $k'$. We then have a stability criterion \begin{eqnarray} \label{eq:SOPT-stability} X_{\sigma-\sigma}({\bf q},0)\left(1 -X_{\sigma-\sigma}({\bf q},0)\right) &\ge& -1 . \end{eqnarray} The linear term on the left-hand side origins from the Hartree approximation. The other term from second-order theory formally enhances the tendency to instability of the Hartree solution, since $X_{\sigma-\sigma}({\bf q},0)<0$. However, the self-consistence renormalizing the one-electron propagator may push the instability to slightly higher values of the interaction strength. \subsection{Iterated perturbation theory} \label{sec:IPT} Iterated perturbation theory is based on the impurity second-order perturbation theory \cite{Georges92,Georges96}. It seems to be up to now the only simple diagrammatic theory capable to describe both the metallic (Fermi liquid) as well as the insulating (atomic limit) phases. This mean-field theory is partially self-consistent with frequencies as dynamical variables. A correction to the Hartree self-energy $\Delta\Sigma_\sigma$ within IPT reads \begin{eqnarray} \label{eq:IPT-self-energy} \Delta\Sigma_\sigma(i\omega_n)&=&-\frac U{\beta }\sum_m {\cal G}_\sigma(i \omega_n+i\nu_m){\cal X}_{-\sigma-\sigma}(i\nu_m) \end{eqnarray} where \begin{eqnarray} \label{eq:G-cal} {\cal G}_\sigma(z)&=&\frac{G_\sigma(z)}{1+G_\sigma(z)\Sigma_\sigma(z)} \end{eqnarray} and ${\cal X}$ is a two-particle bubble defined as in (\ref{eq:2P-bubble}) but with the frequency-dependent (local) propagator ${\cal G}$. We define a new two-particle quantity \begin{eqnarray} \label{eq:IPT-product} \langle{\cal X}_{\sigma\sigma'}(i\nu_m)\rangle&=& \frac U\beta\sum_n \frac 1{1+G_\sigma(i\omega_n)\Sigma_\sigma(i\omega_n)}\ \frac 1{1+G_{\sigma'}(i\omega_n+i\nu_m)\Sigma_{\sigma'} (i\omega_n+i\nu_m)}\nonumber \\ &&\hspace*{-40pt}\times \int d \epsilon\rho(\epsilon) \frac 1{i\omega_n+\mu_\sigma-\Sigma_\sigma(i\omega_n)-\epsilon}\ \frac 1{i\omega_n+i\nu_m+\mu_{\sigma'} -\Sigma_{\sigma'}(i\omega_n+i\nu_m) -\epsilon} \end{eqnarray} needed for the stability criterion of IPT solutions. We easily find that the IPT solution is stable w.r.t spin flips if \begin{eqnarray} \label{eq:IPT-stability} X_{\sigma-\sigma}(0)-{\cal X}_{\sigma-\sigma}(0)\left[\langle {\cal X}_{\sigma-\sigma}(0)\rangle-{\cal X}_{\sigma-\sigma}(0) \right]&\ge&-1 . \end{eqnarray} We see that IPT is generally more stable (to higher values of the coupling constant) than the fully self-consistent perturbation theory from the preceding subsection. It is due to subtraction in the parentheses. This difference stems from a specific dependence of the one-particle propagator ${\cal G}$ on the self-energy. The closer to the impurity case we are (narrow energy band) the smaller the dynamical contribution to the instability of the IPT solution is. In the extreme case of the impurity model, there is no contribution to the instability from second order, i.e. beyond Hartree. Because of the Hartree term there is always a critical interaction at which the IPT solution gets unstable in both the single impurity and the mean-field. This instability, as we already know from the Hartree solution, lies well below the expected Mott-Hubbard metal-insulator transition. It is hence inappropriate to extrapolate a metallic IPT solution to the critical region near the Mott-Hubbard transition and beyond. It seems to be no way of matching consistently the strong-coupling (insulating) and the weak-coupling (metallic) IPT solutions. IPT from mean-field models amounts to a non-self-consistent expansion on finite dimensional lattices. A non-self-consistent expansion is derived within the general scheme of Section~\ref{sec:stability} if the renormalized one-electron propagator is replaced by $G^{(0)-1}=G^{-1}+\Sigma[G]$ where the self-energy is treated as a functional of the full one-electron propagator. All orders of the non-self-consistent expansion get unstable at the same point where the Hartree approximation does. A non-self-consistent expansion does not improve upon stability of the Hartree solution. \subsection{Single-channel approximations} \label{sec:SCh} Dynamical approximations cannot push the instability of the Hartree solution to much higher interaction strengths unless they contain infinite-many two-particle bubbles. Simplest theories with geometric series of two-particle bubbles are single-channel approximations or their extension, the so-called FLEX approximation \cite{Bickers89}. I.~e. we explicitly sum multiple two-particle scatterings on the bare interaction within a single two-particle channel. We derive these approximations within the general scheme of Section~\ref{sec:stability} if we choose $\Lambda^\alpha=U,\ \Lambda^{\alpha'}=0,\ \alpha'\neq\alpha$ \cite{Janis98a}. Since we expect that the instability arises in the electron-hole channel, we explicitly consider only a renormalized RPA (electron-hole channel). A correction to the Hartree self-energy for this approximation is \begin{eqnarray} \label{eq:SCh-self-energy} \Delta\Sigma_\sigma(k)&=&-\frac U{\beta N}\sum_q G_{-\sigma}(k+q)\ \frac{X_{\sigma-\sigma}(q)}{1+X_{\sigma-\sigma}(q)} . \end{eqnarray} The full vertex function $\Gamma$ from the Schwinger-Dyson equation (\ref{eq:sigma-2P}) already contains a geometric series of two-particle bubbles, so we have two nontrivial stability criteria. The one derived from the equation of motion in this approximation reads \begin{eqnarray} \label{eq:SCh-stability1} X_{\sigma-\sigma}({\bf q},0)&\ge&-1 . \end{eqnarray} This is the stability condition from the Hartree approximation with a renormalized one-electron propagator with the self-energy from (\ref{eq:SCh-self-energy}). Condition (\ref{eq:SCh-stability1}) is fulfilled for weakly and moderately correlated systems significantly beyond the instability of the Hartree solution. In the impurity or mean-field cases it is even satisfied up to infinite interaction strength \cite{Hamann69}. A solution fulfilling (\ref{eq:SCh-stability1}) is internally consistent and possesses the necessary analytic properties of the self-energy. It, however, does not guarantee stability of the approximation with respect to external perturbations and to fluctuations beyond this solution. For that purpose we use definition (\ref{eq:2IP-diff}) for the irreducible functions. With the generic notation (\ref{eq:channels}) we obtain a new vertex function in the electron-hole channel \begin{eqnarray} \label{eq:SCh-vertex} \left[\frac{\delta\Sigma}{\delta G}\right]^{eh}_{\sigma-\sigma}[q](k,k') &=& \frac U{1+X_{\sigma-\sigma}(k'-k)}\nonumber\\&& -\frac{U^2}{\beta N} \sum_{k''}\frac{G_{-\sigma}(k+k'+q-k'')}{1+X_{\sigma-\sigma}(k'+q-k'')}\ \frac{G_{\sigma}(k'')}{1+X_{\sigma-\sigma}(k'-k'')} . \end{eqnarray} This vertex function enters the irreducible functions in the other channels, the interaction and the electron-electron ones. We identify in each channel the conserving bosonic variable according to rules (\ref{eq:channels}) and obtain a matrix in $k,k'$ that is to be diagonalized. Unlike second-order we cannot find the lowest real eigenvalue in closed form. We can at least qualitatively assess the lowest eigenvalue in the asymptotic limit of condensation of electron-hole pairs, where the electron-hole two-particle vertex function diverges. First, the electron-electron channel, at least at half filling, is irrelevant and we resort to the electron-hole one. Next, we neglect the convolution of the singular vertices with regular electron-hole bubbles assuming that the integrals smear out the singularities. We hence keep only the first term on the right-hand side of (\ref{eq:SCh-vertex}) as a leading divergent contribution in the asymptotic limit of the critical point. Even then we still have a full $k,k'$ matrix to diagonalize. We further resort to zero temperature with no difference between fermionic and bosonic variables and assume that the lowest eigenvalue can be approximated as in second-order theory. I.~e. we treat the matrix as if it were $k$-independent. This assumption does not hold strictly but only approximately if the influence of the states away from the Fermi surface is not too big. Within this approximation we obtain a new approximate thermodynamic stability criterion in the interaction channel \begin{eqnarray} \label{eq:SCh-stability2} \frac U{\beta N}\sum_{q'} \frac{G_{-\sigma}(({\bf k}_0,0)+q')G_{-\sigma} (({\bf k}_0,0)+q'+({\bf q},0))}{1+X_{\sigma-\sigma}(q')}&\ge&-1 \end{eqnarray} where ${\bf k}_0$ is a suitable vector from the Fermi surface minimizing the expression on the left-hand side. We now have two criteria that both must be fulfilled in order to have a stable phase with correct analytic properties of the Green functions. The former criterion, (\ref{eq:SCh-stability1}), from the equation of motion guarantees internal consistency and analyticity of the approximation. The latter one, (\ref{eq:SCh-stability2}), when fulfilled, implies stability of the approximation with respect to fluctuations beyond the renormalized RPA. When broken, scatterings in the neglected interaction channel become relevant and must self-consistently be interwoven with the electron-hole ones. Simple additions of channels as in the FLEX approximation do not introduce a self-consistence in the two-particle functions and do not improve upon the stability of the single-channel approximations. In low dimensions ($d=2,1$), the left-hand side of (\ref{eq:SCh-stability2}) may even diverge and the single-channel approximations become qualitatively wrong, since the full solution can produce only integrable singularities. Hence FLEX-type approximations are unsuitable for the description of the weak-to-strong-coupling transition. \section{Parquet approximation} \label{sec:parquet} The approximations we have hitherto dealt with were either derived by simplifying the full two-particle vertex function $\Gamma$ in the Schwinger-Dyson equation or some of the Bethe-Salpeter equations were disregarded. The various two-particle functions were hence treated inequivalently and some divergences went lost. The simplest approximation consistent with all the Bethe-Salpeter equations (\ref{eq:Bethe-Salpeter}) is the parquet approximation with the unrenormalized completely irreducible vertex $I[U;G,\Gamma]=U$. We formulate explicitly the stability criteria for this approximation and show how the singularities from the Bethe-Salpeter equations are controlled. To control the singularities that may arise in two-particle functions we diagonalize the Bethe-Salpeter convolutive equations (\ref{eq:Bethe-Salpeter}). We, however, can diagonalize the Bethe-Salpeter equations only separately in each channel and cannot express the solution in closed algebraic form. Diagonalization in the chosen two-particle channel means diagonalization of the appropriate matrix from (\ref{eq:channels}). To formalize this procedure we represent the one- and two-particle functions as operators on a Hilbert space of two-particle (bare) states. Such a Hilbert space can be spanned on the basis vectors $|k,k'>$ parameterized by two one-particle four-momenta. The two-particle vertex functions can be represented as \begin{eqnarray} \label{eq:parquet-vertex} \widehat{X}_{\sigma\sigma'}&=&\sum_{kk'\bar{k}\bar{k}'} \delta(k+k'-\bar{k}-\bar{k}')X_{\sigma\sigma'}(k,k';\bar{k}, \bar{k}') |kk'\rangle\langle \bar{k}'\bar{k}| . \end{eqnarray} We now switch from this channel-independent representation to channel-dependent diagonal representations. We use $q$ for the conserved four-momenta and $Q$ for the eigenvalues of the corresponding matrices in the two-particle channels from (\ref{eq:channels}). We need diagonal representations for the $\alpha$-reducible two-particle functions \begin{eqnarray} \label{eq:parquet-2PR} \widehat{{\cal K}}^\alpha&=&\sum_{q,Q}{\cal K}^\alpha(q,Q)\ |Qq,\alpha \rangle\langle\alpha,qQ| . \end{eqnarray} Note that the eigenvectors $|Qq,\alpha>$ are expressed as a combination of the vectors $|k,k'>$ where $Q$ is a combination of $k$'s. The vector $q$ fixes either difference (electron-hole and interaction channels) or sum (electron-electron channel) of $k',k$. All $|k,k'>$ and $|Qq,\alpha>$ form complete orthogonal sets. For the sake of simplicity we suppressed normalization factors in the sums in (\ref{eq:parquet-vertex}) and (\ref{eq:parquet-2PR}). Representation (\ref{eq:parquet-2PR}) makes the Bethe-Salpeter equations algebraic and we can write their formal solutions as \begin{mathletters}\label{eq:parquet-solution} \begin{eqnarray} \label{eq:parquet-diagonal} {\cal K}^\alpha(q,Q)&=&-\frac{\langle\alpha,qQ|\ \widehat{X}^\alpha\ |Qq, \alpha\rangle}{1+ \langle\alpha,qQ|\ \widehat{Y}^\alpha\ |Qq,\alpha \rangle} \end{eqnarray} where $\widehat{X}$ and $\widehat{Y}$ are two-particle operators. They have the following explicit representation \begin{eqnarray} \label{eq:parquet-X} \langle\alpha,qQ|\ \widehat{X}^\alpha\ |Qq,\alpha\rangle&=& \langle \alpha,qQ|\widehat{U}\widehat{G}^{(2)}\widehat{U}|Qq,\alpha \rangle+\sum_{\alpha'\neq\alpha}\sum_{q',Q'}{\cal K}^{\alpha'}(q',Q') \nonumber\\ &&\hspace*{-110pt} \times\left[\langle\alpha,qQ|\widehat{U}\widehat{G}^{(2)}|Q'q', \alpha'\rangle \langle\alpha',q'Q'|Qq,\alpha\rangle+ \langle\alpha,qQ |Q'q',\alpha'\rangle \langle\alpha',q'Q'|\widehat{G}^{(2)} \widehat{U}|Qq,\alpha\rangle\right] \nonumber\\ &&\hspace*{-110pt} +\sum_{\alpha'\alpha''\neq\alpha}\ \sum_{q'q'',Q'Q''} {\cal K}^{\alpha'}(q',Q') {\cal K}^{\alpha''}(q'',Q'') \langle\alpha,qQ| Q'q',\alpha'\rangle\langle\alpha',q'Q'|\widehat{G} ^{(2)}|Q''q'',\alpha''\rangle\nonumber\\ &&\times\langle\alpha'',q''Q''| Qq,\alpha\rangle , \end{eqnarray} and \begin{eqnarray} \label{eq:parquet-Y} \langle\alpha,qQ|\ \widehat{Y}^\alpha\ |Qq,\alpha\rangle&=&\langle \alpha,qQ|\widehat{U}\widehat{G}^{(2)}|Qq,\alpha \rangle+\sum_{\alpha'\neq\alpha}\sum_{q',Q'}{\cal K}^{\alpha'}(q',Q') \nonumber\\ && \times\langle\alpha,qQ|\widehat{U}\widehat{G}^{(2)}|Q'q',\alpha' \rangle\langle\alpha',q'Q'|\widehat{G}^{(2)}| Qq,\alpha\rangle . \end{eqnarray} \end{mathletters} We used abbreviations for the free two-particle propagation \begin{mathletters}\label{eq:parquet-aux} \begin{eqnarray} \label{eq:parquet-G2} \widehat{G}^{(2)}_{\sigma\sigma'}&=&\sum_{kk'}G_\sigma(k)G_{\sigma'}(k') |kk'\rangle\langle k'k| \end{eqnarray} and for the operator of the bare vertex \begin{eqnarray} \label{eq:parquet-U} \widehat{U}_{\sigma\sigma'}&=&U\delta_{\sigma',-\sigma}\sum_{kk'\bar{k} \bar{k}'} \delta(k+k'-\bar{k}-\bar{k}')\ |kk'\rangle\langle \bar{k}' \bar{k}| . \end{eqnarray} \end{mathletters} The parquet approximation has become a set of closed equations for ${\cal K}^\alpha(q,Q)$ for the given one-electron propagators. The most difficult task in solving parquet equations (\ref{eq:parquet-solution}), (\ref{eq:parquet-aux}) is the diagonalization in the separate channels, i.~e. finding the bases $|Qq,\alpha>$ represented in the bare states $|kk'>$. They are determined from (\ref{eq:channels}) and (\ref{eq:2IP-def}) so that the parquet equations remain implicit and must be iterated. However, parquet equations (\ref{eq:parquet-solution}) allow us to control the behavior of the denominators, i.~e. the appearance of singularities. To close the parquet approximation we have to rewrite the Schwinger-Dyson equation with the diagonal form of the vertex function. It reads \begin{eqnarray} \label{eq:parquet-sigma} \Delta\Sigma_{\sigma}(k)&=&-\sum_{k'}G_{-\sigma}(k')\left[\langle k'k| \widehat{U}\widehat{G}^{(2)}_{\sigma-\sigma}\widehat{U}|kk'\rangle +\sum_{\alpha}\sum_{q,Q}{\cal K}^{\alpha}_{\sigma-\sigma}(q,Q) \right.\nonumber\\ &&\hspace*{80pt} \times \langle k'k|Qq,\alpha\rangle\langle\alpha,qQ|\widehat{U}\widehat{G}^{(2)} _{\sigma-\sigma}|kk'\rangle\Bigg] . \end{eqnarray} The parquet approximation is completely determined by numbers $\Sigma_\sigma(k)$, ${\cal K}^\alpha_{\sigma\sigma'}(q,Q)$ and $<\alpha,qQ|kk'>$. The internal stability criteria for the parquet approximation in the diagonal representation are \begin{eqnarray} \label{eq:parquet-stability} 1+ \langle\alpha,({\bf q},0)({\bf Q},0)|\ \widehat{Y}^\alpha\ |({\bf Q},0) ({\bf q},0),\alpha\rangle&\ge&0 . \end{eqnarray} We see that all the two-particle functions can contain only integrable singularities and that if a singularity arises in one channel it cannot cause a havoc in the other channels, since the integral kernels are integrable functions. There are also stability criteria for the parquet approximation with respect to external perturbations beyond this approximate equilibrium state. Since possible instabilities must be integrable we do not expect a qualitatively different behavior. A qualitative change could be caused by new bound states with more than two particles that could not be decomposed into simpler two-particle states. It is physically plausible to assume nonexistence of such states, at least all the existing approaches do that. It is a tremendous task to solve the complete parquet approximation. It can be completed only in a very limited special situations \cite{Bickers98}. However, the general scheme how to solve the parquet equations helps us to identify the principal features any simplification should not loose. It is a two-particle self-consistence allowing only for integrable singularities and an access to the diagonal or algebraic form of the solution of the Bethe-Salpeter equations enabling to control possible singularities. \section{Simplifications in the parquet approach} \label{sec:simplifications} Parquet diagrams differ significantly from the simpler single-channel approximations with multiple two-particle scatterings. The difference manifests itself in the effective interaction in the two-particle scattering processes. The bare interaction remains unrenormalized in the single-channel approximations. In the parquet equations, it is replaced by renormalized vertex functions representing a dynamically screened inter-particle potential. A qualitative difference between a solution to the full parquet equations and to the single-channel approximations becomes evident at the two-particle criticality. The vertex functions from one or more channels get critical and sharply peaked around the Fermi energy. The exchange between the scattered particles becomes strongly energy dependent and cannot be approximated by a static effective interaction as in the single-channel theories. The Bethe-Salpeter nonlinear integral equations are extremely complicated with an intricate analytic structure of the solution. Each two-particle function entering the parquet equations contains initially three independent (complex) variables. Neither of them can simply be neglected, since the variables are interconnected due to a mixing of different inequivalent channels. Even the diagonal form of the parquet equations with only two relevant variables remains implicit and does not allow for a solution in closed form. That is why a nonperturbative solution to the parquet equations has not yet been found. From this reason the parquet diagrams are relatively little used in condensed matter although being in existence more than thirty years. The only hope that the parquet approximation may become useful for quantitative assessments of the effects of strong electron correlations is that one succeeds in simplifying the full parquet algebra to an analytically manageable form. As already discussed, the parquet approach is expected to produce qualitatively new results at intermediate and strong coupling with critical two-particle functions where weak-coupling approximations get unstable. In simplifying the parquet equations we will demand asymptotic accuracy at the two-particle criticality, particularly for condensation of electron-hole pairs leading at half filling to a metal-insulator transition. We hence retain only the diagrams potentially making the effective mass of the electrons divergent at the metal-insulator transition. \subsection{Two-Variable Two-Particle Functions} It was recently argued that at least at half filling only two of the three topologically inequivalent channels are relevant in the critical region of the metal-insulator transition. Only multiple scatterings in the electron-hole and the interaction channels contain divergent diagrams at the critical point \cite{Janis98a}. This simplification to two channels itself, called dipole approximation, does not reduce the number of variables in the two-particle functions. It only simplifies the algebra of the parquet approximation. To lower the number of the relevant variables we use the fact that the singularities in the two-particle functions in the parquet approximation must be integrable. It means that when integrating over the variable in which the singularity arises we obtain a finite result bounded by a constant of order one. Our approximation consists in neglecting all finite (nondiverging) contributions at the critical point. This must be done at the level of the diverging two-particle functions. As a first step we reduce the number of the relevant variables in the two-particle functions to two. This is achieved when we neglect mixing of the reducible functions from different channels. We replace the full two-particle function $\Gamma$ on the right-hand side of the Bethe-Salpeter equation in the $\alpha$-channel by $U+{\cal K}^\alpha$, i.~e. by a sum of the completely irreducible vertex and the reducible function in that channel. Such a replacement does not change the denominator of the diagonal representation of the solution to the Bethe-Salpeter equations and hence the universal critical behavior of the function ${\cal K}^\alpha$ remains unaltered. The bare interaction is a constant and consequently the solution does not depend on the outgoing (incoming) variables. Further on, the triplet functions can be eliminated from this approximation, since $\Lambda_{\sigma,\sigma}$ gets irrelevant. If we denote $\Lambda^v_{\sigma,-\sigma}=U+{\cal K}^{eh}_{\sigma,-\sigma}$ and $\Lambda^h_{\sigma,-\sigma}=U+{\cal K}^U_{\sigma,-\sigma}$, the parquet equations reduce to \begin{mathletters}\label{eq:I} \begin{eqnarray} \label{eq:parq2a} \Lambda^v_{\sigma,-\sigma}(k;q)&=&U\nonumber\\ &&\hspace*{-10pt} -\frac 1{\beta{\cal N}}\sum_{q'}\Lambda^h_{\sigma,-\sigma}(k;q')G_\sigma(k+q') G_{-\sigma} (k+q'+q) \Lambda^v_{\sigma,-\sigma}(k+q';q)\ , \\[4pt] \label{eq:parq2b} \Lambda^h_{\sigma,-\sigma}(k;q)&=&U+\frac 1{\beta^2{\cal N}^2}\sum_{q',q''} \Lambda^v_{\sigma,-\sigma}(k;q')G_{-\sigma}(k+q') G_{-\sigma}(k+q'+q) \nonumber\\ &&\hspace*{-10pt} \times \Lambda^v_{-\sigma,\sigma}(k+q';q''-q') G_\sigma (k+q'') G_\sigma (k+q''+q)\Lambda^h_{\sigma,-\sigma}(k+q'';q)\ . \end{eqnarray}\end{mathletters} Because of the broken symmetry between the incoming and outgoing variables in the two-particle functions we must be careful in simplifying the equation for the self-energy. To be consistent we use a symmetrized formula with the bare interaction at the incoming and outgoing momenta. We obtain \begin{eqnarray} \label{eq:Sigma2} \Delta\Sigma_\sigma(k)&=&-\frac U{2\beta^2{\cal N}^2}\sum_{k',q} G_\sigma(k+q) G_{-\sigma}(k') G_{-\sigma}(k'+q)\left[\Lambda^h_{\sigma,-\sigma}(k+q; -q)\right. \nonumber\\ &&\left. + \Lambda^h_{-\sigma,\sigma}(k';q)+ \Lambda^v_{\sigma,-\sigma}(k+q;k' -k)+\Lambda^v_{-\sigma,\sigma}(k';k-k') -2U\right]\ . \end{eqnarray} The reduced two-channel parquet approximation with two-variable two-particle functions reproduces in leading order the critical behavior of the parquet approximation (with two channels). The universal quantities derived from the divergent functions of the parquet equations are reproduced by equations (\ref{eq:I})-(\ref{eq:Sigma2}). Only nonuniversal quantities such as the critical interaction do not come out in this simplification as in the full parquet theory. We achieved a simplification of the full parquet algebra without loosing leading critical behavior of the solution, but the resulting equations retain a nonlinear convolutive character. It is still too elaborate to reach nonperturbative solutions or to diagonalize them. We must find a further reduction of complexity of the parquet equations to make them useful for an affordable quantitative analysis. \subsection{One-Variable Two-Particle Functions} It is the bosonic variable $q$ in the two-particle functions $\Lambda^h,\Lambda^v$ that is relevant at the critical point and in which the singularity arises. The fermionic variable $k$ in both functions ``labels'' the eigenvalues of the integral kernel in equations (\ref{eq:parq2a}), (\ref{eq:parq2b}). The minimal eigenvalue governs the critical behavior at low temperatures. If we are interested only in the asymptotic critical behavior we can approximate the actual minimal eigenvalue by a value at the lowest-lying fermionic four-momentum. By such an approximation we neglect mixing of the fermionic four-momenta in the Bethe-Salpeter equations and $k$-dependence in the two-particle functions. This approach corresponds to a low-energy expansion used by Hamann \cite{Hamann69} by assessing the strong-coupling limit of the renormalized RPA of Suhl in the single-impurity problem. We assume that the approximate minimal eigenvalue dominates and qualitatively determines the relevant physics of the critical point. The suggested simplification is possible only at zero temperature, where the difference between the fermionic and bosonic Matsubara frequencies vanishes. We hence put $T=0$, which is the most interesting case for the metal-insulator transition. Using the above ansatz, neglecting $k$-dependence in $\Lambda^h,\Lambda^v$, and putting $k=({\bf k}_0,0)$ in the one-electron propagators we reduce the parquet equations to a manageable form. We introduce new functions \begin{mathletters}\label{eq:vertex} \begin{eqnarray} \label{eq:Gam} \Gamma_{\sigma}(q)&=&\frac 1{\beta{\cal N}}\sum_{q'} \Lambda^h_{\sigma,-\sigma}(q') G_{\sigma}(({\bf k}_0,0)+q')G_{-\sigma} (({\bf k}_0,0)+q'+q)\\ \label{eq:K} {\cal K}_{\sigma}(q)&=& \frac 1{\beta{\cal N}}\sum_{q'} \Lambda^v_{-\sigma,\sigma}(q') G_{\sigma}(({\bf k}_0,0)+q')G_{\sigma} (({\bf k}_0,0)+q'+q)\ , \end{eqnarray}\end{mathletters} with the aid of which we obtain for the two-particle vertex functions \begin{mathletters} \begin{eqnarray} \label{eq:Iv} \Lambda^v_{\sigma,-\sigma}(q)&=&\frac U{1+\Gamma_{\sigma}(q)}\ ,\\[4pt] \label{eq:Ih} \Lambda^h_{\sigma,-\sigma}(q)&=& \frac U{1-{\cal K} _{-\sigma}(q){\cal K}_{\sigma}(q)} \ . \end{eqnarray}\end{mathletters} The introduced fermionic momentum $({\bf k}_0,0)$ from the Fermi surface is an adjustable parameter chosen to maximize the tendency to instability of this approximation. The above equations have the complexity of the single-channel approximations, but contain renormalized vertex functions in the two-particle scatterings. It is always the renormalized vertex function that enters the denominator of equations (\ref{eq:I}) via the functions $\Gamma$ and ${\cal K}$ and the critical behavior is properly renormalized. It is easy to derive the corresponding equation for the self-energy. It reads \begin{eqnarray} \label{eq:sigma1} \Delta\Sigma_\sigma(k)&=&-\frac 1{2\beta{\cal N}}\sum_q\left[G_\sigma(k+q) X_{-\sigma,-\sigma}(q) \left(\Lambda^h_{-\sigma,\sigma}(q)+\Lambda^h _{\sigma,-\sigma}(-q)-U\right)\right.\nonumber\\ &&\hspace*{20pt}\left. +G_{-\sigma}(k+q)X_{\sigma,-\sigma}(q)\left( \Lambda^v_{\sigma,-\sigma}(q)+I^v_{-\sigma,\sigma}(-q) -U\right)\right] \end{eqnarray} where $X_{\sigma,\sigma'}$ are two-particle bubbles with the renormalized one-electron propagators. Equations (\ref{eq:vertex})-(\ref{eq:sigma1}) replace the original parquet approximation. They fully determine the generating two- and one-particle functions. All thermodynamic and spectral quantities can be calculated from them. Since we used a number of approximate steps in the derivation of the reduced parquet algebra, any new quantity must first be formulated within the full parquet approximation. The Luttinger-Ward functional generating all thermodynamic quantities in the parquet approximation and simplifications thereof is constructed in the Appendix. Simplifications are introduced in the exact formulas via the generating one- and two-particle functions. Having a simplified and manageable set of parquet-type equations we can finally formulate internal stability criteria for them. They are derived from the general ones and resemble the stability criteria from the single-channel approximations. They read \begin{eqnarray} \label{eq:simplified-stability} 1+{\cal K}_{\sigma}({\bf q},0)\ge 0 &,\hspace{10mm}& 1+\Gamma_{\sigma}( {\bf q},0)\ge 0 . \end{eqnarray} Unlike the single-channel approximations, stability criteria (\ref{eq:simplified-stability}) contain renormalized dynamical two-particle vertex functions that are determined self-consistently from the simplified parquet equations. The functions $\Gamma_\sigma$ and ${\cal K}_{\sigma}$ depend implicitly on the momentum ${\bf k}_0$ from the Fermi surface emerging due to consistency of our simplification. The vector ${\bf k}_0$ is chosen so that to minimize the left-hand side of (\ref{eq:simplified-stability}) The vector then determines the character of the symmetry breaking ($s$ vs. $d$ wave etc.) first appearing in the system. \section{Dynamical mean-field approximation and metal-insulator transition} \label{sec:DMFT} We apply the dipole (simplified parquet) approximation to effective impurity models (single impurity or dynamical mean-field) at intermediate and strong coupling. The fluctuations in space do not contribute to the dynamics and the four-momenta from the general formulation of the parquet approximation collapse to frequencies. This simplification enables one to perform explicitly analytic continuation from the Matsubara to the real frequencies using contour integrals. Realizing that the reduced equations hold for $T=0$ and introducing $\zeta,z$ for the bosonic, fermionic complex frequencies, respectively, we can write for the two-particle functions \begin{mathletters} \label{eq:2P-cont} \begin{equation} \label{eq:K-cont} {\cal K}_\sigma(\zeta)=-U\int_{-\infty}^0\frac{d\omega}\pi \left[G_\sigma(\omega+\zeta)\mbox{Im}\frac{G_\sigma(\omega_+)} {1+\Gamma_{-\sigma}(\omega_+)} + \frac{G_\sigma(\omega-\zeta)} {1+\Gamma_{-\sigma}(\omega-\zeta)}\mbox{Im}G_\sigma(\omega_+)\right], \end{equation} \begin{eqnarray} \label{eq:Gam-cont} \Gamma_\sigma(\zeta)&=&-U\int_{-\infty}^0\frac{d\omega}\pi \left[G_{-\sigma}(\omega+\zeta)\mbox{Im}\frac{G_\sigma(\omega_+)} {1-{\cal K}_\sigma(\omega_+){\cal K}_{-\sigma}(\omega_+)} \right. \nonumber\\[4pt] &&\hspace*{80pt}\left. +\frac{G_\sigma(\omega-\zeta)}{1-{\cal K}_\sigma(\omega-\zeta){\cal K}_{-\sigma}(\omega-\zeta)} \mbox{Im}G_{-\sigma}(\omega_+)\right] \end{eqnarray}\end{mathletters} and analogously for the one-particle self-energy correction to the Hartree term \begin{eqnarray} \label{eq:Sigma1} \Delta \Sigma_\sigma(z)&=&\int_{-\infty}^0\frac{d\omega}\pi\left\{ G_\sigma(\omega+z) \mbox{Im}\left[X_{-\sigma,-\sigma}(\omega_+)\left( \Lambda^h_{-\sigma,\sigma}(\omega_+) -U\right)\right]\right.\nonumber\\ &&\hspace*{-50pt}\left. +X_{-\sigma,-\sigma} (\omega-z)\left(\Lambda ^h_{-\sigma,\sigma}(\omega_+) -U\right)\mbox{Im}G_\sigma(\omega_+) +G_{-\sigma}(\omega+z)\mbox{Im}\left[X_{\sigma,-\sigma}(\omega_+) \Lambda^v_{\sigma,-\sigma}(\omega_+)\right] \right.\nonumber \\ &&\left. +X_{\sigma,-\sigma}(\omega-z)\Lambda^v_{\sigma,-\sigma}(\omega-z) \mbox{Im}G_{-\sigma}(\omega_+)\right\} \end{eqnarray} where we denoted $\omega_+=\omega+i0^+$. Equations (\ref{eq:2P-cont}) and (\ref{eq:Sigma1}) can be solved numerically. We use the one-electron propagators from the Bethe lattice in $d=\infty$ with a finite half-bandwidth $w=1$. We as well put $B=0$, since the Mott-Hubbard metal-insulator transition is expected there and both channels should show the same divergence at this critical point. We have three levels of self-consistence in the parquet equations. We have to determine self-consistently the Hartree parameters (spin-dependent particle densities), then the dynamical self-energy correction, and finally the two-particle vertex functions (cf. Appendix, (\ref{eq:Omega})). It appears that we have to perform the self-consistences sequentially from top to bottom. First the Hartree must be self-consistently determined, second the self-energy correction, and finally the vertex functions. Only in this order we can secure convergence of the iteration scheme and analytic properties of the converged and intermediate solutions. First step in our iteration scheme for the two-particle functions reduces to self-consistent second order. It becomes unstable at around $U\approx w$ and gets spin polarized at intermediate coupling. Iterations to the FLEX approximation with two (electron-hole and interaction) channels converge very quickly at weak and intermediate coupling. The converged solution never gets spin-polarized. However, starting with $U\approx w$ the solution becomes unstable with respect to external perturbations, spin-flips, and condition (\ref{eq:SCh-stability2}) is violated. Fig.~\ref{fig:FLEX-instab} shows the real part of the vertex function $\Gamma$ for the FLEX with the first next iteration towards the parquet solution. A stable solution demands that both functions at $\omega=0$ lie above the value $-1$. \begin{figure} \epsfig{figure=flex-instab2.ps,height=12cm, angle=-90} \caption{\label{fig:FLEX-instab} Vertex function $\Gamma$ calculated within the FLEX (solid line) and first iteration beyond FLEX (dashed line). Their values at zero frequency decide about stability of the approximation ($\mbox{Re}\Gamma(0)>-1$).} \end{figure} The iterations for the reduced dipole approximation converge at weak and intermediate coupling for $U\le 2w$. We experienced problems with convergence for higher values of the interaction strength \cite{note2}. Due to logarithmic singularities we had insufficient resolution close to the Fermi energy. We used a linear distribution of the frequency mesh points with $\Delta\omega=0.005w$, since it is convenient for evaluating the convolutions. We need all the energy scales to reach a solution fulfilling necessary sum rules, but as the metal-insulator transition is approached, a broad spectrum of energy scales is to be covered. A more sophisticated method for dealing with the different energy scales in equations (\ref{eq:2P-cont}), such as numerical renormalization group, has to be applied if we want to continue the approximation closer to the metal-insulator transition. Although we were unable to iterate the solutions for interactions approaching the metal-insulator transition, we nevertheless can demonstrate a qualitative difference in the behavior of the FLEX and the parquet approximations. Apart from the fact that the FLEX approximation becomes unstable, the main qualitative difference is evident for small frequencies. Fig.~\ref{fig:FLEX-comp2} shows the difference in the frequency behavior of the real part of the vertex function $\Gamma$ at $U=2w$. A tendency to forming of a narrow quasiparticle peak around the Fermi energy is evident in the parquet approximation. The parquet approximation is much closer to the instability point $\Gamma(0)=-1$ than the FLEX solution. However, due to the self-consistence in the parquet approach, the solution remains stable. \begin{figure} \epsfig{figure=comp_2P_u2.ps,height=12cm, angle=-90} \caption{\label{fig:FLEX-comp2}Comparison of the two-particle vertex calculated at $U=2w$ within the FLEX and the simplified parquet approximations. The inset shows the detailed structure around the Fermi energy.} \end{figure} Fig.~\ref{fig:FLEX-comp-ImG} and Fig.~\ref{fig:FLEX-comp-ImS} show the same for the imaginary part of the one-electron propagator and the self-energy. A tendency to the expected insulating solution and splitting of the quasiparticle central band at low frequencies can again be clearly recognized. However, the low-energy dynamics in the parquet approximation is different from the unstable IPT solution \cite{Georges96}. The central quasi-particle peak is much more narrow, however much less separated from the excited states. No pronounced quasigap is yet present. The high frequency behavior of the density of states should be dominated by satellite peaks. Emerging of shoulders for the expected lower and upper Hubbard bands can be seen in the parquet solution (but not in FLEX). Also the satellite bands (the high-energy region) are less pronounced than in the IPT and numerical renormalization group \cite{Bulla98} or Monte Carlo simulations \cite{Jarrell92,Georges96}. Hence the separation of low- and high-energy scales seems to go much slower in the diagrammatic approaches than in numerical simulations. It can be explained by the fact that the spin-symmetric atomic limit with a discrete energy spectrum is not easily reproducible in the parquet construction with a continuous spectrum of energies. It is the half-bandwidth $w$ that fixes the energy scale and alike the Bethe-ansatz solution it cannot simply be limited to zero. No tendency to a divergence in the self-energy is observed, as seen in the numerical renormalization group \cite{Bulla98}. It is impossible to produce a singular self-energy in the metallic phase in the Schwinger-Dyson equation with integrable two-particle functions. \begin{figure} \epsfig{figure=comp_ImG_u2.ps,height=12cm, angle=-90} \caption{\label{fig:FLEX-comp-ImG}Comparison of the imaginary part of the one-electron propagator from the FLEX and the simplified parquet approximations. The inset shows a tendency towards the formation of a central quasiparticle peak in the parquet approximation.} \end{figure} \begin{figure} \epsfig{figure=comp_ImS_u2.ps,height=12cm, angle=-90} \caption{\label{fig:FLEX-comp-ImS}Comparison of the imaginary part of the self-energy with the layout as in Fig.~\ref{fig:FLEX-comp-ImG}.} \end{figure} The parquet approximation has two vertex functions that are to show the same divergence. However, their dynamical behavior beyond the Fermi energy is different as shown in Fig.~\ref{fig:parq-2P}. It is the irreducible vertex function from the interaction channel ${\cal K}$ that has a stronger tendency towards instability. The values of the two vertex functions at the Fermi level coincide within the numerical resolution and lie very close to $-1$, within the precision of $10^{-9}$. \begin{figure} \epsfig{figure=2P_u2.ps,height=12cm, angle=-90} \caption{\label{fig:parq-2P}Real parts of the two vertex functions $\Gamma$ and ${\cal K}$ calculated with the simplified parquet approximation at $U=2w$. Although these functions coincide at the Fermi energy their dynamical behavior is different.} \end{figure} The critical point of the metal-insulator transition is not directly accessible with the present numerical solution of the simplified parquet approximation. However, the low frequency behavior of the divergent functions can be assessed in leading order analytically. At the critical point of the metal-insulator transition both vertex functions, $U/(1+\Gamma(\omega))$ and $U/(1+{\cal K}(\omega))$, diverge at the Fermi energy $\omega=0$. Consequently, the derivatives of the real part of the self-energy (effective mass) and the derivatives of the imaginary parts of the vertex functions diverge. It is straightforward to obtain in leading order \begin{mathletters}\label{eq:sing} \begin{eqnarray} \label{eq:K-Gam-sing} \mbox{Im}\ {\cal K}'_\sigma(i0^+)&\doteq& -\frac U\pi\ \frac{\left [ \mbox{Im}\ G_\sigma(i0^+)\right]^2}{1 +\Gamma_{-\sigma}(0)}\ , \hspace{10pt}\mbox{Im}\ \Gamma'_\sigma(i0^+)\doteq -\frac U\pi\ \frac{\mbox{Im}\ G_{\sigma}(i0^+)\mbox{Im}\ G_{-\sigma}(i0^+)} {1-{\cal K}_{-\sigma}(0){\cal K}_{\sigma}(0)}\ ,\\[4pt] \label{eq:Sigma-sing} \mbox{Re}\ \Sigma'_\sigma(0)&\doteq&-\frac U\pi\ \left[ \mbox{Im}\ G_{-\sigma}(i0^+)\frac{X_{\sigma-\sigma}(0)}{1+\Gamma_\sigma(0)} +\mbox{Im}\ G_{\sigma}(i0^+)\frac{X_{-\sigma-\sigma}(0)}{1-{\cal K}_{-\sigma}(0){\cal K}_{\sigma}(0)} \right] \end{eqnarray}\end{mathletters} where the small (Kondo) energy scale is $\Delta=w(1+\Gamma(0))$ or equivalently $\Delta=w(1+{\cal K}(0))$. We see from (\ref{eq:sing}) that the divergence of the effective mass is bound to a divergence in the electron-hole dominated irreducible vertex functions. It is important to stress that a singularity in one channel causes a nonanalyticity in the other channel. Hence suppressing any of the channels with electron-hole loops in the effective impurity problems means inappropriate treatment of the metal-insulator (Kondo) critical region. If the Schwinger-Dyson equation of motion is fulfilled the effective mass cannot diverge, or a sharp metal-insulator transition cannot appear without a singularity in the electron-hole (local) vertex functions. Fig.~\ref{fig:parq-div} shows for $U=2w$ the three functions the slopes of which should be infinite at the transition point. The least sharp slope displays the vertex function $\Gamma$ as can be expected from the more enhanced tendency to instability in ${\cal K}$ than in $\Gamma$. The spin-symmetric situation is the most difficult to analyze because of the degeneracy in the spin space (singlet and triplet scatterings are indistinguishable). If we switch on an external magnetic field we remove the spin symmetry and the density of states will no longer be pinned at the Fermi level. The value of ${\cal K}_{\sigma}(0){\cal K}_{-\sigma}(0)$ decreases and we push the solution away from the criticality of the metal-insulator transition. Going from weak to strong coupling at a fixed field, we gradually increase the absolute value of $\Gamma_\sigma(0)$ as long as we reach a critical point (for sufficiently small fields) $\Gamma_\sigma(0)=-1$ at the boundary to the fully spin-polarized (Hartree) state $U_{s}$. The value of ${\cal K}_{\sigma}(0){\cal K}_{-\sigma}(0)$ first increases with interaction, reaches a maximum at intermediate coupling and then, due to the Hartree band-splitting terms in the self-energy, decreases to zero at the boundary to the fully spin-polarized state. There is no indication for a Mott-Hubbard metal-insulator transition in an external magnetic field. An extrapolation from finite fields to $B=0$ suggests a hypothesis that the metal-insulator critical point is shifted to infinity as in the single impurity case. However, neither finite fields do allow for a stable numerical solution with linearly distributed mesh points at strong coupling to support the hypothesis quantitatively. A detailed quantitative analysis of the Hubbard model in an external magnetic field will be presented elsewhere. \begin{figure} \epsfig{figure=divergence_u2.ps,height=12cm, angle=-90} \caption{\label{fig:parq-div} Three functions with divergent slopes at the metal-insulator transition from (\ref{eq:sing}) at $U=2w$.} \end{figure} \section{Discussion and conclusions} \label{sec:conclusions} We demonstrated in this paper the importance of a proper formulation of approximate theories in order to determine the range of their applicability. We showed that a careful stability analysis is needed in intermediate or strong-coupling regimes due to instabilities, poles that may arise in two-particle functions. We must have a reliable description of two-particle functions to be able to decide about stability of solutions. We suggested that the parquet approach formulated here as an extension of the Baym-Kadanoff construction with properly chosen external disturbances offers the desirable framework for a consistence and stability analysis of approximate solutions. Under a physically plausible argument that singularities in two-particle functions are due to poles in the Bethe-Salpeter equations, the parquet approach leads to a qualitatively correct asymptotic critical behavior at transition points. It introduces a new two-particle self-consistence causing that two-particle functions can contain only integrable divergences as in the exact theory. We found a complete set of criteria for local instabilities from the poles in the Bethe-Salpeter equations. We attached to each singularity in two-particle functions an order parameter breaking a symmetry of the underlying Hamiltonian. The parameters are new self-energies and are either of density type, such as magnetization or have anomalous character and do not conserve either spin (transverse magnetic order) or charge (superconductivity). Since any approximate self-consistent theory breaks some relations between various Green functions, we always have two sets of two-particle functions and stability criteria. The two-particle functions determined from the Schwinger-Dyson equation of motion are used for internal stability and consistency of the theory and determine whether spectral functions have the desired analytical properties. The two-particle functions defined thermodynamically from the Luttinger-Ward functional via functional derivatives are used to determine external stability w.r.t. external disturbances or dynamical fluctuations. The two different two-particle functions must produce qualitatively the same phase diagram and spectral properties in a trustworthy theory. Approximations in the Baym scheme are introduced for the full two-particle vertex function in the Schwinger-Dyson equation. Since the vertex function may get singular, approximations to the Schwinger-Dyson equation can make the resulting theory unreliable at two-particle criticality. The presented parquet approach does not touch the Schwinger-Dyson and Bethe-Salpeter equations of motion and introduces approximations only in the input to these equations, namely in the completely irreducible two-particle vertex. This is a new qualitative change in describing critical behavior, since unlike the full vertex function, the completely irreducible vertex is not expected to cause new singularities to appear. Diagrammatically, it can only be broken into two separate parts by cutting at least three one-particle propagators. It reduces in the parquet approximation to the bare interaction. A lowest-order correction to it is plotted in Fig.~\ref{fig:beyond-parquet}. \begin{figure} \epsfig{figure=parqfig2.eps,height=6cm} \caption{\label{fig:beyond-parquet} A lowest-order contribution to the completely two-particle irreducible vertex beyond the parquet approximation. The bare interaction is replaced by the full two-particle vertex from the parquet solution. Double primed variables are summation indices.} \end{figure} Corrections to the completely irreducible vertex beyond the parquet approximation consist of integrals over potentially singular functions and do not generate new divergences. These corrections may, however, change the critical exponents \cite{Bickers92}. However, an expansion beyond the parquet approximation must not be interchanged with a weak-coupling loop expansion. The parquet approximation can be viewed upon as a sophisticated ``mean-field'' theory with dynamical vertex corrections for strongly correlated systems. Strongly correlated electrons do not allow for simpler static mean-field theories, since two-particle criticality may realistically be described only if we do not loose integrability of divergences and stability in all two-particle channels. All simpler approaches loose this essential property of the exact theory. The most difficult problem with the parquet approach is that even the simplest approximation does not allow for explicit nonperturbative solutions in closed form. Hence we need further simplifications to obtain a workable approximate scheme. We proposed such a scheme and tested it on effective impurity models with frequencies as dynamical variables. The main idea of our simplification is to abandon all irrelevant variables that do not influence the existence and quality of singularities in two-particle functions. We reduced the number of relevant two-particle channels at half filling to two and kept only the variables at which the divergence in the appropriate channel arises. The resulting simplified theory is comparable in complexity with the single-channel (FLEX) approximations, but behaves in a qualitatively different manner at intermediate and strong coupling. We showed explicitly on the Hubbard and single-impurity Anderson models in an external magnetic field that all simpler approximations get unstable at intermediate coupling and fail to describe the transition from weak to strong coupling. Only the parquet approximation remained stable on the metallic side of the Mott-Hubbard metal-insulator transition. Although analytically stable, the simplified parquet approximation with linearly distributed integration mesh points became numerically unstable when used in the critical region of the Mott-Hubbard metal-insulator transition. This instability is of purely numerical character and does not follow from the quality of the parquet approximation itself. A picture from the parquet theory for the transition from weak to strong coupling in the mean-field solution of the Hubbard model has not yet been completed, but the presented analysis strongly suggests in agreement with \cite{Kehrein98}, that there cannot be a sharp, second-order Mott-Hubbard transition for finite coupling at zero temperature. First, we have not found a sharp transition when an external magnetic field is applied and the only divergence in the two-particle spin-flip function was found at the boundary to the fully spin-polarized solution. Second, if there were a sharp transition in the paramagnetic, spin-symmetric phase, then inevitably two local two-particle functions from the electron-hole and the interaction channels had to diverge at the transition point. It follows from our stability analysis that then a local order should set in. However, local two-particle functions are sums of momentum-dependent two-particle functions over the Brillouin zone, even in infinite dimensions. If such a sum diverges, it must be at least one momentum ${\bf q}_0$ at which the paramagnetic solution gets unstable before the Mott-Hubbard transition is reached. Hence, a long-range order will conceal the Mott-Hubbard metal-insulator critical point. We conclude that Fermi liquid can break down only if a long-range order sets in. Based on the experience and understanding of the origin of singularities at intermediate and strong coupling, we can conclude that only theories with dynamical vertex renormalizations are capable to describe the transition from weak to strong coupling qualitatively correctly. The parquet approximation with an adequate numerical technique to iterate or diagonalize the resulting nonlinear equations offers a suitable analytic-numerical tool for this purpose. Especially quantum criticality is the situation where only the parquet approach is able to give a complete picture of the critical behavior and symmetry-breaking order parameters. What is to be done in particular cases is to find appropriate simplifications making the parquet theory viable but retaining the essential physics. To summarize, we have presented a general parquet construction treating one- and two-particle Green functions on the same footing within a single self-consistent renormalization scheme. We showed that instabilities are induced by divergences in two-particle functions to which symmetry-breaking order parameters are always attributed, either normal of density type or anomalous. Further on, an instability in an approximate theory need not definitely indicate a transition to a long-range order. A singularity in one two-particle channel induces singular vertex corrections in the other channels that tend to suppress the long-range order. Even if a divergence appears only in one channel, as in an external magnetic field, a singularity in the other channels is generated and at least one another channel is needed to introduce a two-particle self-consistence and to keep the solution stable. The parquet approach is always to be used where quantum phase transitions are present and stability criteria in at least two channels may be broken. It is the low-temperature limit in low spatial dimensions ($d=0,1,2$) and in intermediate- and strong-coupling regimes in three and higher dimensions. There, the parquet approximation is the simplest consistent and qualitatively correct description. The work was supported by grant No. 202/98/1290 of the Grant Agency of the Czech Republic. \begin{appendix} \setcounter{equation}{0} \renewcommand{\thesection}{} \renewcommand{\theequation}{\Alph{section}.\arabic{equation}} \section{} To derive the generating thermodynamic functional for the parquet approximation we proceed as generally outlined in \cite{Dominicis62} and explicitly evaluated for the parquet approximation in \cite{Janis98b}. I.~e. we introduce pairwise Legendre conjugate variables for the irreducible and reducible two-particle vertex functions. We then integrate the Bethe-Salpeter equations (\ref{eq:Bethe-Salpeter}) so that \begin{eqnarray} \label{eq:Phi-dif} {\cal K}^\alpha=\frac{\delta\Phi_\Lambda}{\delta\Lambda^{\alpha}} \ , &\hspace{1cm}& \Lambda^\alpha= \frac{\delta\Phi_K}{\delta{\cal K}^ {\alpha}}\ . \end{eqnarray} We use the notation introduced in Sec.~\ref{sec:stability} and particularly (\ref{eq:channels}) and write the functionals as traces over the active variables in each channel \begin{eqnarray} \label{eq:Phi-U} \Phi^{U}_\Lambda&=&\frac 1{2\beta{\cal N}}\sum_{q}\mbox{Tr}_{\sigma k} \left\{\left[\Lambda^{U}G^{(2)}\right]^U[q]-\frac 12\left[\Lambda^{U} G^{(2)}\right]^U[q]\star\left[\Lambda^{U}G^{(2)}\right]^U[q]\right. \nonumber \\ &&\hspace*{60pt} \left. -\ln_\star\left(1+\left[\Lambda^{U}G^{(2)}\right][q]\right) \right\} \ , \end{eqnarray} \begin{eqnarray} \label{eq:Phi-eh} \Phi^{eh}_\Lambda&=&\frac 1{2\beta{\cal N}}\sum_{\sigma\sigma' q}\mbox{Tr}_{k} \left\{\left[\Lambda^{eh}G^{(2)}\right]^{eh}_{\sigma\sigma'}[q] -\frac 12\left[\Lambda^{eh} G^{(2)}\right]^{eh}_{\sigma\sigma'}[q]\bullet \left[\Lambda^{eh}G^{(2)}\right]^{eh}_{\sigma\sigma'}[q]\right. \nonumber \\ &&\hspace*{60pt} \left. -\ln_\bullet\left(1+\left[\Lambda^{eh}G^{(2)}\right]^{eh} _{\sigma\sigma'}[q]\right)\right\} \ , \end{eqnarray} \begin{eqnarray} \label{eq:Phi-ee} \Phi^{ee}_\Lambda&=&\frac 1{2\beta{\cal N}}\sum_{\sigma\sigma' q}\mbox{Tr}_{k} \left\{\left[\Lambda^{ee}G^{(2)}\right]^{ee}_{\sigma\sigma'}[q] -\frac 12\left[\Lambda^{ee} G^{(2)}\right]^{ee}_{\sigma\sigma'}[q]\circ \left[\Lambda^{ee}G^{(2)}\right]^{ee}_{\sigma\sigma'}[q]\right. \nonumber \\ &&\hspace*{60pt} \left. -\ln_\circ\left(1+\left[\Lambda^{ee}G^{(2)}\right]^{ee}_{\sigma \sigma'}[q]\right)\right\} \ \end{eqnarray} where $G^{(2)}$ is the free two-particle propagator defined in (\ref{eq:parquet-G2}). Note that the trace in the interaction channel includes spin as an active variable. The logarithms with subscript denote the appropriate matrix multiplication used in their power series expansions. Integration of the other set of two-particle equation of the parquet approximation leads to a functional \begin{eqnarray} \label{eq:Phi-K} \Phi_K&=&\frac 1{2\beta{\cal N}}\sum_{\sigma\sigma' q}\mbox{Tr}_{k}\left\{ \left[\left(U+{\cal K}^{ee}\right)G^{(2)}\right]^{eh}_{\sigma\sigma'}[q] \bullet\left[{\cal K}^{eh}G^{(2)}\right]^{eh}_{\sigma\sigma'}[q] +\left[UG^{(2)}\right]^{ee}_{\sigma\sigma'}[q]\circ\left[{\cal K}^{ee} G^{(2)}\right]^{ee}_{\sigma\sigma'}[q]\right\} \nonumber\\ && + \frac 1{2\beta{\cal N}}\sum_{q}\mbox{Tr}_{\sigma k}\left\{\left[ \left(U+{\cal K}^{eh}+{\cal K}^{ee}\right)G^{(2)}\right]^U[q]\star \left[{\cal K}^U G^{(2)}\right]^U[q]\right\} \ . \end{eqnarray} Be aware that the operator of the bare interaction acts only between different spins, i.~e. contains $\delta_{\sigma,-\sigma}$. The functionals $\Phi^\alpha_\Lambda$ are exact and only the functional $\Phi_K$ is approximate, since only it depends on the completely irreducible vertex being replaced here by the bare interaction. The Luttinger-Ward generating functional is a sum of $\Phi^\alpha_\Lambda$ and $\Phi_K$ completed with second-order term and products of the Legendre conjugate two-particle functions $\Lambda$ and ${\cal K}$ to keep the variation of the Luttinger-Ward functional w.r.t. two-particle functions zero. We obtain \begin{eqnarray} \label{eq:Phi-full} \Phi[G;\Lambda,{\cal K};U]&=&\frac 1{2\beta{\cal N}}\sum_{q}\mbox{Tr} _{\sigma k}\left\{\left[\Lambda^U{G}^{(2)}\right]^U[q]\star\left[ {\cal K}^U G^{(2)}\right]^U[q]-\frac 12\left[UG^{(2)}\right]^U[q] \star\left[UG^{(2)}\right]^U[q]\right\}\nonumber\\ &&\hspace*{-50pt}+\frac 1{2\beta{\cal N}}\sum_{\sigma\sigma' q}\mbox{Tr}_{k} \left\{\left[\Lambda^{eh}G^{(2)}\right]^{eh}_{\sigma\sigma'}\bullet \left[{\cal K}^{eh}G^{(2)}\right]^{eh}_{\sigma\sigma'}[q]+\left[ \Lambda^{ee}G^{(2)}\right]^{ee}_{\sigma\sigma'}[q]\circ\left[ \Lambda^{ee}G^{(2)}\right]^{ee}_{\sigma\sigma'}[q]\right\} \nonumber\\[4pt] &&-\Phi_K\left[G;{\cal K};U\right]-\Phi^{eh}_\Lambda\left[G;\Lambda^{eh} \right]-\Phi^{ee}_\Lambda\left[G;\Lambda^{ee}\right]-\Phi^{U}_\Lambda \left[G;\Lambda^{U}\right] \end{eqnarray} where we introduced in each functional its explicit dependence on the variational one- and two-particle functions and the bare interaction $U$. The grand potential generating the complete thermodynamics of the parquet approximation is constructed from the Luttinger-Ward functional and a free-electron term with the Hartree contribution \cite{Baym62}. We hence add also the particle densities as additional variational parameters and obtain for a fixed $\mu_\sigma=\mu+\sigma B$ \begin{eqnarray} \label{eq:Omega} \frac 1{{\cal N}}\Omega [n_{\uparrow},n_{\downarrow};\Sigma ,G;\Lambda, {\cal K}]&=&-\frac 1{\beta {\cal N}}\sum_{\sigma n,{\bf k}}e^{i\omega_n0^{+}}\Big\{ \ln \left[ i\omega _n+\mu _\sigma -\epsilon({\bf k}) -Un_{-\sigma }-\Sigma_\sigma ({\bf k},i\omega_n)\right] \nonumber\\[4pt] & & \hspace*{10pt} +G_\sigma ({\bf k},i\omega_n)\Sigma _\sigma ({\bf k},i\omega_n)\Big\} -Un_{\uparrow}n_{\downarrow} +\Phi [G;\Lambda, {\cal K};U] \ . \end{eqnarray} Here $n_{\uparrow},n_{\downarrow};\Sigma ,G;\Lambda,{\cal K}$ are independent variables and complex functions. Their physical values are chosen from stationarity of the grand potential (\ref{eq:Omega}) with respect to variations of its independent variables/functions. There are three pairs of Legendre conjugate variational variables in (\ref{eq:Omega}). The Hartree parameters $n_\uparrow$ and $n_\downarrow$, the one-electron functions $\Sigma_\sigma(k)$ and $G_\sigma(k)$, and finally the two-particle irreducible and reducible functions $\Lambda^\alpha_{\sigma \sigma'}(k,k';q)$ and ${\cal K}^\alpha_{\sigma\sigma'}(k,k';q)$, respectively. \end{appendix}
\section{Introduction} \noindent Quantum mechanics of constrained systems is experiencing a new wave of interest connected with the recent progress in semiconductor physics: nowadays experimentalists are able to investigate the behavior of electrons in structures of various shapes, at times rather elaborated. The small size, extreme material purity, and its crystallic structure make it possible to derive basic properties of these systems in a crude but useful model in which the electron is considered as a free particle (with an effective mass) whose motion is constrained to a prescribed subset of $\Real^d$ with $d=2,3$, possibly in presence of external fields. On the theoretical side, this inspires questions about relations between spectral and scattering properties of such systems and the underlying geometry and topology. A class of systems which attracted a particular attention are {\em quantum waveguides}, \ie\/ tubular regions supporting a Schr\"odinger particle. It is known that a deviation from the straight tube can induce existence of bound states and resonances in scattering, vortices in probability current, \etc, be it bending \cite{DE1, DEM, DES, ES1, GJ}), protrusion or a similar local deformation \cite{AS, BGRS, EV1}, waveguide coupling by crossing \cite{SRW}, or by one or several lateral windows \cite{ESTV, EV2, EV3} (the related bibliography is rather extensive; the quoted papers contain many more references). In this paper we are going to discuss a system closely related to the last named one. It supposes again a double waveguide; however, the coupling between the two parallel ducts will entail now a tunneling through a thin semitransparent barrier rather than a window in a hard wall separating them --- \cf~Figure~\ref{schm}. To get a solvable model we describe the barrier by a $\delta$ potential whose coupling strength may vary longitudinally: the Hamiltonian can be then formally written as \begin{equation} \label{formal} H_{\alpha}= -\Delta_{\Omega}+\alpha(x)\delta(y) \end{equation} with the barrier supported by the $x$--axis, where $\Omega:=\Real\times(-d_2,d_1)$ is the double--guide strip. \begin{figure}[!htb] \begin{center} \input fig01 \end{center} \caption{Double waveguide with a $\delta$ barrier.}\label{schm} \end{figure} There are several motivations to investigate a leaky--barrier waveguide pair. First of all, it is a generalization in a sense of earlier results, because the pierced--hard--wall case of Ref.~\cite{ESTV} corresponds to $\alpha=0$ in the window and $\alpha=\infty$ otherwise. Recall that the latter can serve to describe an actual quantum--wire coupler --- see, \eg,~\cite{HTW, Ku} --- and such a model will certainly become more realistic if the tunneling through the barrier of a doped semiconductor material separating the two guides is taken into account. At the same time, the Hamiltonian~(\ref{formal}) covers for various $\alpha$ a wide variety of situations. On the mathematical side, the $\delta$ potential of~(\ref{formal}) can be treated more easily than the hard--wall barrier, since two operators with different functions $\alpha$ have the same form domain. To illustrate the difference, one can compare the variational proof of existence of bound states in Thm~\ref{existence} below with the analogous argument of Ref.~\cite{ESTV}. A deeper application of the quadratic--form perturbations allows us to construct the Birman--Schwinger theory for the waveguide systems in question, in particular, to derive the weak--coupling behaviour of the bound states. This will be done in a subsequent paper~\cite{EK}. Let us describe briefly the contents of the paper. In the next section we shall describe the model and deduce its spectrum in the ``unperturbed'', \ie\/ translationally invariant case. In Section~\ref{Sec.Existence} we demonstrate that a local change of the coupling parameter will cause the existence of bound states provided it is negative in the mean. To ilustrate the spectral and also scattering properties we shall discuss then in detail the example in which the barrier function is of a ``rectangular well'' shape. In the final section we will show how the situation modifies if the semitransparent barrier is placed at the surface of a cylinder threaded by a homogeneous magnetic field. \setcounter{equation}{0} \section{Preliminaries} \subsection{The Hamiltonian} Let $\Omega:=\Real\times{\cal O}$ with ${\mathcal O}:={\cal O}_{2}\cup{\cal O}_{1}:= (-d_{2},0)\cup(0,d_{1})$ be the configuration space, \ie, the part of $\Real^2\,$ occupied by the waveguide. Passing to the rational units, $\hbar=2m=1$, we may identify the particle Hamiltonian $H_{\alpha}$ with the Laplace operator away of the waveguide boundary and the barrier. To give meaning to the formal expression (\ref{formal}) one has to specify the boundary conditions. At the outer edges we assume the Dirichlet condition, \begin{equation} \label{Dirichlet} \psi(x,-d_{2})=\psi(x,d_{1})=0, \end{equation} while the barrier is transversally the usual $\delta$ potential defined conventionally as \begin{equation} \label{barrier} \psi(x,0+)=\psi(x,0-)=:\psi(x,0), \quad \psi_{y}(x,0+)-\psi_{y}(x,0-)=\alpha(x) \psi(x,0) \end{equation} for any $x\in\Real$ --- \cf~\cite[Sec~.I.3]{AGH} --- where the subscript denotes partial derivative with respect to $y$. The Hamiltonian domain is then \begin{equation}\label{Hamiltonian} D(H_{\alpha}):= \left\{ \psi\in\sobii(\Omega)| \;\: \psi\quad {\rm satisfies}\, (\ref{Dirichlet}) \;{\rm and}\; (\ref{barrier})\; \right\}, \end{equation} where the function $\alpha:\Real\rightarrow\Real$, assumed to be piecewise continuous, determines the shape of the barrier and represents the $x$-dependent coupling ``constant'' of the interaction. For the sake of simplicity we shall exclude the above mentioned case of a Dirichlet barrier, $\alpha(x)=\infty$ at a subset of $\Real$. In that case all the operators $H_{\alpha}$ have the same form domain, and the associated quadratic form is obtained by a simple integration by parts: \begin{eqnarray} q_{\alpha}[\psi] &:=& \int_{\Real\times{\bar{\cal O}}}|\nabla\psi|^2(x,y)dx dy+ \int_{\Real}\alpha(x)|\psi(x,0)|^2 dx, \label{form} \\ \nonumber \\ D(q_{\alpha}) &:=& \{\psi\in\sobi\left(\Real\times(-d_{2},d_{1})\right)|\ \forall x\in\Real:\psi(x,-d_{2})=\psi(x,d_{1})=0\}.\label{formdomain} \end{eqnarray} The form (\ref{form}) is obviously symmetric and it is not difficult to check that it is closed and thus indeed associated with $H_{\alpha}$. Hereafter we adopt the notation of \cite{ESTV}: \mbox{$d:=\max\{d_{1},d_{2}\}$}, \mbox{$D:=d_{1}+d_{2}$}, and \[\nu:=\frac{\min\{d_{1},d_{2}\}}{\max\{d_{1},d_{2}\}}.\] Without loss of generality we may assume that $d_{2}\leq d_{1}=d$. \subsection{The unperturbed system} If $\alpha(x)=\alpha$ is a constant function, we can solve the Schr\"odinger equation $H_{\alpha}\psi=k^2\psi$ by separation of variables. To get the transverse eigenfunctions we have to match smoothly the solutions in the two ducts, \mbox{$C_{2}\sin \ell(y+d_{2})$} and \mbox{$C_{1}\sin \ell(y-d_{1})$}, chosen to satisfy the condition (\ref{Dirichlet}). If $\ell d_{1}, \ell d_{2}$ are not multiples of $\pi$ we get thus the following condition on eigenvalues of the transverse part of the Hamiltonian: \begin{equation}\label{spc} -\alpha=\ell\,(\cot \ell d_{1}+\cot \ell d_{2}). \end{equation} \vspace{2mm} \begin{rem}\label{rem1}{\rm If $d_{1},d_{2}$ are rationally related the Schr\"odinger equation can be also solved by $\ell=\frac{\pi n p}{d_{1}}=\frac{\pi n q}{d_{2}},\; n\in\Nat\setminus\{0\}$. However, such wave functions are zero at $y=0$, and therefore independent of $\alpha$. In this sense they represent a trivial part of the problem. A prime example is the symmetric waveguide pair, $d_{1}\!=\!d_{2}$, where this observation concerns every solution antisymmetric w.r.t. $y\!=\!0$. It is reasonable to concentrate on the nontrivial part only. If $\nu\equiv\frac{d_{2}}{d_{1}}=\frac{p}{q}$, we denote by ${\cal G}_{\nu}$ the subspace in $\sii(-d_{2},d_{1})$ spanned by the solutions of (\ref{spc}). Putting then ${\cal H}_{\nu}:=\sii(\Real)\otimes{\cal G}_{\nu}^{\bot}$, we shall restrict our attention to the operator $H_{\alpha}\upharpoonright\,{\cal H}_{\nu}$; for the sake of simplicity we shall denote the restriction by the symbol $H_{\alpha}$ again. The trivial part is absent, of course, if $\nu$ is irrational.} \end{rem} \noindent From the spectral condition (\ref{spc}) we get a sequence of eigenvalues (in the natural ascending order) of (the nontrivial part of) the transverse operator; we denote it as $\{\nu_{n} (\alpha)\}_{n=1}^{\infty}$. The corresponding eigenfunctions are \begin{equation}\label{eigenfunctions} \chi_{n}(y;\alpha)= (-1)^j N_{n}\, \sin \sqrt{\nu_{n}}d_{j}\, \sin \sqrt{\nu_{n}}\left(y\!+\!(-1)^jd_{j}\right) \end{equation} for $y\in{\cal O}_{j},\; j=1,2$, where $N_{n}$ is the normalization factor chosen in such a way that $\chi_n$ would be a unit vector in $\sii(-d_{2},d_{1})$, \ie \begin{equation}\label{norms} N_{n}^{2}= \frac{2\sqrt{\nu_{n}}} {\sqrt{\nu_{n}}d_{1}\sin^{2}\sqrt{\nu_{n}}d_{2}+ \sqrt{\nu_{n}}d_{2}\sin^{2}\sqrt{\nu_{n}}d_{1}- \sin\sqrt{\nu_{n}}d_{1}\sin\sqrt{\nu_{n}}d_{2}\sin\sqrt{\nu_{n}}D}\,. \end{equation} Furthermore, the Green's function of the Hamiltonian (\ref{formal}) can be written down explicitly: \begin{equation}\label{green0} G_{\alpha}(x,y,x',y';k)=\sum_{n=1}^{\infty}\, \frac{i}{2 k_{n}}\, e^{i k_{n}|x-x'|}\ \chi_{n}(y;\alpha)\,\bar\chi_{n}(y';\alpha)\,, \end{equation} where the efective momentum in the $n$-th transverse mode is $k_{n}:=\sqrt{k^2-\nu_{n}(\alpha)}$. Elementary properties of the transverse eigenvalues follow from the the spectral condition~(\ref{spc}) by means of the implicit-function theorem; we collect them in the lemma below. \begin{lemma}\label{approxbound} (a) Let $\{m_{i}\}_{i=0}^{\infty}$ be the sequence obtained from the set $\Nat\,\cup\,\nu^{-1}\Nat$ by natural ordering. Then $ \frac{\pi}{2 d}(n\!-\!1) \leq\frac{\pi}{d}m_{n-1} <\sqrt{\nu_{n}} <\frac{\pi}{d}m_{n} \leq\frac{\pi}{d}n\,$ holds for all $n\in\Nat\setminus\{0,1\}$. \\ [1mm] (b) The function $\alpha\mapsto\nu_{n}(\alpha)$ is strictly increasing and continuous for all $n\in\Nat\setminus\{0\}$. \end{lemma} \setcounter{equation}{0} \section{Existence of bound states}\label{Sec.Existence} Depending on the choice of $\alpha$, the operators (\ref{formal}) offer a variety of spectral types. In this paper we shall concentrate on the situation when the barrier describes a local perturbation of the system with separating variables considered above. The locality is at that understood as a decay of the function $\alpha$; in other words, we shall assume that $\lim_{|x|\to\infty}\alpha(x)=\alpha_0$. It is important that the limiting value $\alpha_0$ is the same at both directions. In such a case, it is easy to localize the essential spectrum. One employs a simple bracketing argument similar to that of~\cite[Sec.~II]{ESTV}) squeezing $H_{\alpha}$ between a pair of operators with Dirichlet and Neumann conditions on segments perpendicular to the $x$-axis placed to both sides of the centre. By the minimax principle only the tails of the estimating operators contribute to their essential spectra; since the ``cuts" can be chosen arbitrarily far we obtain $\sigma_{ess}(H_{\alpha})= [\nu_{1}(\alpha_{0}),\infty)$. Less trivial is the existence of discrete spectrum. It is known that any ``window" in the impenetrable barrier induces a bound state. This fact was established first for sufficiently wide windows \cite{Po}, later an independent and more general proof proof was given \cite{ESTV} with no lower bound on the window width. The present case is more complicated because the effective coupling strength $\alpha-\alpha_0$ can be sign--changing. We shall show that it is sufficient if it is negative in the mean creating thus a locally average stronger tunelling between the two channels: \begin{thm}\label{existence} Assume that {\em (i)} $\;\alpha\!-\!\alpha_{0}\in L_{loc}^1 (\Real)$, {\em (ii)} $\;\alpha(x)\!-\!\alpha_{0}={\cal O}(|x|^{-1-\varepsilon})$ for some $\varepsilon>0$ as $|x|\to\infty$. If $\;\int_{\Real}(\alpha(x)\!-\!\alpha_{0})dx<0$, then $H_{\alpha}$ has at least one isolated eigenvalue below its essential spectrum. \end{thm} \PF We use a variational argument whose idea comes back to ~\cite{GJ}; see also~\cite{DE1, RB}, and~\cite[Sec.~III]{ESTV}) for a coupled waveguide system. First of all, the assumption (ii) tells us that $\lim_{|x|\to\infty}|x|^{1+\varepsilon} (\alpha(x)\!-\!\alpha_{0})=0$, \ie, to any $\delta>0$ there is $a_{\delta}>1$ such that \begin{equation}\label{delta} |x|>a_{\delta} \;\Rightarrow\; |\alpha(x)\!-\!\alpha_{0}| <\frac{\delta}{|x|^{1+\varepsilon}}. \end{equation} It is useful to introduce a shifted energy form: for an arbitrary $\Psi\in D(q_{\alpha})$ we put % \begin{equation}\label{Q} Q_{\alpha}[\Psi]:=q_{\alpha}[\Psi]-\nu_{1}(\alpha_{0})\|\Psi\|_{2}^{2}; \end{equation} since the essential spectrum of $H_{\alpha}$ starts at $\nu_{1}(\alpha_{0})$, we have to find a trial function $\Psi$ such that $Q_{\alpha}[\Psi]$ is negative. We obtain it by a suitable modification of the function $\Psi_0(x,y):= \chi_{1}(y;\alpha_{0})$ which formally annuls (\ref{Q}) for $\alpha=\alpha_0$ but does not belong to $L^2$. The trial function has to decay; in order to make the positive contribution from its tails to the kinetic energy small, we employ an exterior scaling. We choose an interval ${\cal A}:=[-a,a]$ for some $a>1$ and a function $\varphi\in{\cal S}(\Real)$ in such a way that $\varphi(x)\le 1$ and $\varphi(x)=1$ on ${\cal A}$. Then we can define the family $\{\varphi_{\sigma}:\sigma\in\Real\}$ by a scaling exterior to ${\cal A}$: \begin{equation}\label{scaling} \varphi_{\sigma}(x):=\left\{ \begin{array}{lcl} \varphi(x) &\quad\mbox{if}\quad& |x|\leq a\\ &&\\ \varphi(\pm a+\sigma(x\mp a)) &\quad\mbox{if}\quad& \pm x>a \end{array} \right. \end{equation} By construction, $|\varphi_{\sigma}(x)|\leq 1$ holds for all $x\in\Real$. The sought trial function will be chosen in the form $\Psi(x,y):= \varphi_{\sigma}(x)\chi_{1}(y;\alpha_{0})$. We employ the relations $\|\dot{\varphi}_{\sigma}\|_{2}^{2}= \sigma\|\dot{\varphi}\|_{2}^{2}$, and \begin{eqnarray*} q_{\alpha}[\Psi]&=&q_{\alpha_{0}}[\Psi]+ \int_{\Real}(\alpha(x)\!-\!\alpha_{0})|\Psi(x,0)|^2 dx,\\ q_{\alpha_{0}}[\Psi]&=&\|\dot{\varphi}_{\sigma}\|_{2}^{2} +\nu_{1}(\alpha_{0})\|\varphi_{\sigma}\|_{2}^{2}\,, \end{eqnarray*} the last one of which is obtained by tedious but straightforward calculation. This yields \begin{equation}\label{Q=} Q_{\alpha}[\Psi]=\sigma\|\dot{\varphi}\|_{2}^{2} +|\chi_{1}(0;\alpha_{0})|^2\int_{\Real} (\alpha(x)\!-\!\alpha_{0})|\varphi_{\sigma}(x)|^2 dx\,. \end{equation} We split now the integration region into two mutually disjoint parts, ${\cal A}$ and $\Real\setminus{\cal A}$. Using (\ref{delta}) together with the above mentioned bound on $\varphi_{\sigma}$ we arrive at the estimate \begin{equation}\label{Q<} Q_{\alpha}[\Psi]<\sigma\|\dot{\varphi}\|_{2}^{2} +\frac{4\,\delta\,|\chi_{1}(0;\alpha_{0})|^2}{\varepsilon a^{\varepsilon}} +|\chi_{1}(0;\alpha_{0})|^2\int_{\Real}(\alpha(x)\!-\!\alpha_{0})\,dx\,. \end{equation} By assumption we have $\int_{\Real}(\alpha(x)\!-\!\alpha_{0}) \,dx<0$ and since $\chi_{1}(0;\alpha_{0})$ is nonzero, the last term is negative; it is then enough to choose $\delta$ and $\sigma$ sufficiently small to make $Q_{\alpha}[\Psi]$ negative.\qed \begin{rem}\label{WeakRem}{\rm A case of particular interest concerns weakly coupled Hamiltonian of the type (\ref{formal}), \ie\/ the situation when $\alpha$ differs from $\alpha_0$ only slightly. In that case one can develop a Birman-Schwinger analysis in order to derive the perturbative expansion of the ground state energy in terms of a parameter measuring the ``smallness" of $\alpha\!-\!\alpha_0$. This will be done in a separate paper \cite{EK}; here we just borrow a result for a further use in this work. There are different ways in which $\alpha\!-\!\alpha_0$ can be small. Suppose that the support of the perturbation shrinks, \ie\/ introduce $\alpha_{\sigma}(x):= \alpha(x/\sigma)$ with the scaling parameter $\sigma\in(0,1]$ and consider the limit $\sigma\to 0+$. We have the following result~\cite{EK}: \begin{thm}\label{WeakThm} Suppose that $\alpha\!-\!\alpha_{0}$ is non-zero and belongs to $L^{1+\varepsilon}(\Real,dx) \cap\si(\Real,|x|^2 dx)$ for some $\varepsilon>0$. Then $H_{\alpha_{\sigma}}$ has for small $\sigma$ at most one simple eigenvalue $E(\sigma)<\nu_{1}(\alpha_0)$, and this happens if and only if $\int_{\Real}(\alpha(x)\!-\!\alpha_{0})\,dx\leq 0$. If this condition holds the following expansion is valid \begin{eqnarray}\label{expansion} \lefteqn{\sqrt{\nu_{1}-E(\sigma)} = -\frac{\sigma}{2}\,|\chi_{1}(0;\alpha_0)|^2 \int_{\Real}(\alpha(x)\!-\!\alpha_{0})\,dx} \nonumber\\ && +{\sigma^2\over 4}\, |\chi_{1}(0;\alpha_0)|^2\sum_{n=2}^{\infty}|\chi_{n}(0;\alpha_0)|^2 \int_{{\Real}^2}(\alpha(x)\!-\!\alpha_{0})\, \frac{e^{-\sigma\sqrt{\nu_{n}-\nu_{1}}|x-x'|}} {\sqrt{\nu_{n}-\nu_{1}}}\, (\alpha(x')\!-\!\alpha_{0})\,dx\,dx\nonumber\\ && +{\cal O}(\sigma^3). \end{eqnarray} \end{thm}} \end{rem} \setcounter{equation}{0} \section{A ``rectangular well'' example}\label{Sec.Example} To illustrate the above result and to analyze the behaviour of coupled waveguides in more details we shall now investigate an example. We choose the barrier function $\alpha$ so that the corresponding Schr\"odinger equation can be solved numerically; this happens if $\alpha$ is a step-like function which makes it possible to employ the mode-matching method. The simplest nontrivial case concerns a ``rectangular well" of a width $2a>0$, \begin{displaymath} \alpha(x):=\left\{ \begin{array}{lcl} \alpha_{1} & \mbox{if} & |x|<a\\ &&\\ \alpha_{0} & \mbox{if} & |x|\leq a \end{array} \right. \end{displaymath} for some $\alpha_{1},\alpha_{0}\in\Real$. In view of Theorem~\ref{existence} this waveguide system has bound states if and only if $\alpha_1< \alpha_0$. In particular, one expects that in the case when $\alpha_{1}=0$ and $\alpha_{0}$ is large positive the spectral properties will be similar to those of the impenetrable barrier situation studied in~\cite{ESTV}. On the other hand, the mode-matching method allows us to treat on same footing the scattering processes in our waveguide. Then there is no need to impose the above condition, because the ``barrier" situation, $\alpha_1> \alpha_0$ is expected to exhibit nontrivial scattering behaviour as well. Henceforth, we shall denote the transverse eigenvalues in the two regions as $\nu_n^{s}:=\nu_n(\alpha_s)$, $s:=0,1$, $\;n\in\Nat\setminus\{0\}$. In view of the natural mirror symmetry with respect to the $y$-axis we may consider separately the symmetric and antisymmetric solutions, \ie\/ to analyze the halfstrip with the Neumann or Dirichlet boundary condition at $x=0$, respectively. For the sake of simplicity we shall also restrict our attention to the case $\;\min\{\alpha_0,\alpha_1\} >\alpha_m:= -(d_{1}^{-1}+d_{2}^{-1})$, when the lowest transverse eigenvalue is positive everywhere in the waveguide. The considerations presented below remain valid even without this assumption; one has just to replace the trigonometric ground-state eigenfunction for hyperbolic which makes the formulae cumbersome. \subsection{Bound states} \label{Sec.existence} Let us first derive an estimate which allows to localize roughly the eigenvalues. It is based on a bracketing argument similar to that used to specify the essential spectrum at the beginning of Section~\ref{Sec.Existence}. The Hamiltonian can be squeezed between a pair of operators, $H_{\alpha}^{(N)}\leq H_{\alpha}\leq H_{\alpha}^{(D)}$, with additional Dirichlet/Neumann ``cuts" at segments perpendicular to the waveguide axis, $x=\pm a$. The spectra of the estimating operators can be easily found and sought estimate comes from the eigenvalues of the middle part situated below $\nu^0_1$ in combination with the minimax principle. In particular, we find that the number $N$ of isolated eigenvalues satisfies the bounds $$ N_{D}+1\geq N\geq N_{D}:=\left[\frac{2 a}{\pi}\sqrt{\nu_{1}^{0}-\nu_{1}^{1}}\right]\,, $$ where $[\cdot]$ denotes the entire part; this complements Theorem~\ref{existence}. Furthermore, the $n$-th eigenvalue $E_{n}$ of $H_{\alpha}$ is estimated by \begin{equation}\label{estE} \nu_{1}^{1}+\left(\frac{(n-1)\pi}{2 a}\right)^2 \leq E_n \leq \nu_{1}^{1}+\left(\frac{n\pi}{2 a}\right)^2, \end{equation} while the critical halfwidth $a_n$ at which the $n$-th eigenvalue emerges from the continuum satisfies the bounds \begin{equation}\label{estg} \frac{(n-1)\pi}{2\sqrt{\nu_1^{0}-\nu_1^{1}}} \leq a_n \leq \frac{n\pi}{2\sqrt{\nu_1^{0}-\nu_1^{1}}}. \end{equation} After this preliminary, let us pass to the mode-matching method. We start with the simpler case when the waveguide exhibits a mirror symmetry w.r.t. the $x$-axis, \ie\/ $d_1=d_2=d$. \subsubsection{The symmetric case} If $\nu=1$, the Hamiltonian decouples into an orthogonal sum of the even and the odd parts, the spectrum of the latter being clearly trivial --- \cf~Remark~\ref{rem1}. The two symmetries allow us to restrict ourselves to the part of $\Omega$ in the first quadrant, with Neumann or Dirichlet condition in the segment $(0,d)$ of the $y$-axis, and take the transverse eigenvalues determined by the spectral conditions \begin{eqnarray*} -\alpha_0 &=& 2 \ell\cot\ell d \qquad\mbox{if}\quad x\geq a\\ -\alpha_1 &=& 2 \ell\cot\ell d \qquad\mbox{if}\quad 0\leq x<a. \end{eqnarray*} The corresponding transverse eigenfunctions are \begin{eqnarray} \chi_{n} &:=& -N_n^{0}\sin\sqrt{\nu_n^{0}}(y\!-\!d) \qquad\mbox{if}\quad x\geq a,\nonumber \\ \phi_{n} &:=& -N_n^{1}\sin\sqrt{\nu_n^{1}}(y\!-\!d) \qquad\mbox{if}\quad 0\leq x<a, \end{eqnarray} where $N_n^{s}$ is a normalization factor chosen to make $\chi_n,\phi_n$ unit vectors in $\sii(0,d)$, \ie \begin{equation} (N_n^{s})^2=\frac{4\sqrt{\nu_n^{s}}} {2\sqrt{\nu_n^{s}}d-\sin 2\sqrt{\nu_n^{s}}d}\ . \end{equation} The overlap integrals of elements of the two bases are easily seen to be \begin{equation}\label{overlap} (\chi_m,\phi_n)=\frac{N_m^0 N_n^1}{\nu_m^0-\nu_n^1} \left(\sqrt{\nu_n^1}\sin\sqrt{\nu_m^0}d\,\cos\sqrt{\nu_n^1}d -\sqrt{\nu_m^0}\sin\sqrt{\nu_n^1}d\,cos\sqrt{\nu_m^0}\right). \end{equation} A natural Ansatz for the solution of an energy $E\in[\nu_1^1,\nu_1^0)$ is \begin{equation}\label{Ansatz} \begin{array}{rll} \Psi_{s/a}(x,y)&=\sum\limits_{n=1}^{\infty}b_n^{s/a} e^{-q_n(x-a)}\chi_n(y) &\qquad\mbox{for}\quad x\geq a\\ &&\\ \Psi_{s/a}(x,y)&=\sum\limits_{n=1}^{\infty}a_n^{s/a} \left\{ \begin{array}{c} \frac{\cosh p_n x}{\cosh p_n a}\\ \\ \frac{\sinh p_n x}{\sinh p_n a} \end{array} \right\} \phi_n(y) &\qquad\mbox{for}\quad 0\leq x<a \end{array} \end{equation} where the subscripts (we will omit them for the most part) $s,a$ distinguish the symmetric and antisymmetric case, respectively. The longitudinal momenta are defined by \begin{displaymath} q_n:=\sqrt{\nu_n^0-E}\ ,\qquad p_n:=\sqrt{\nu_n^1-E}\ . \end{displaymath} As an element of the domain~(\ref{Hamiltonian}), the function $\Psi$ should be continuous together with its normal derivative at the segment dividing the two regions, $x=a$. Using the orthonormality of $\{\chi_n\}$ we get from the requirement of continuity \begin{equation}\label{contRequir} b_m=\sum_{n=1}^{\infty}a_n(\chi_m,\phi_n)\,. \end{equation} In the same way, the normal-derivative continuity at $x=a$ yields \begin{equation}\label{normal-derRequir} b_m q_m+\sum_{n=1}^{\infty}a_n p_n \left\{ \begin{array}{c} \tanh\\ \coth \end{array} \right\} (p_n a) (\chi_m,\phi_n)=0\,. \end{equation} Substituting from~(\ref{contRequir}) to~(\ref{normal-derRequir}), we can rewrite it as an operator equation \begin{equation}\label{C.a=0} \boldmath{C a}=\boldmath{0}\,, \end{equation} where \begin{equation} C_{mn}:=\left(q_m+p_n \left\{ \begin{array}{c} \tanh\\ \coth \end{array} \right\} (p_n a)\right)(\chi_m,\phi_n) \end{equation} with the overlap integrals given by~(\ref{overlap}). It is straightforward to compute the norms of the functions~(\ref{Ansatz}); since $n^{-1}q_n$ an $n^{-1}p_n$ tend to $\frac{\pi}{d}$ as $n\to\infty$ (see~Lemma~\ref{approxbound}~1.), the square integrability of $\Psi$ requires the sequences $\{a_n\}$ and $\{b_n\}$ to belong to the space $\ell^2(n^{-1})$. To make sure that the equation~(\ref{C.a=0}) makes sense, it is enough to notice that if $\Psi$ is an eigenvector of $H_{\alpha}$, it must belong to the domain of any integer power of this operator. It is easy to check that \begin{equation} \forall i\in\Nat\!\setminus\!\{0\}:\quad \Psi\in D(H_{\alpha}^i)\Leftrightarrow \{a_n\},\{b_n\}\in\ell^2(n^{4 i-1})\,; \end{equation} hence the sought sequences should belong to $\ell^2(n^r)$ for all $r\geq -1$, \ie\ both sequences have a faster than powerlike decay. This fact also justifies {\em a posteriori\/} the interchange of summation and differentiation we have made in the matching procedure. Furthermore, one can use it to check the existence of a convergent series of cut-off approximants to the solutions in the same way as in~\cite[Sec.~IV.1]{ESTV}. \begin{rem}[an alternative method]{\rm We can use the orthonormality of $\{\phi_n\}$ instead of $\{\chi_n\}$ and express $\{a_n\}$ in analogy to~(\ref{contRequir}), and then substitute it into~(\ref{normal-derRequir}). We find that the coefficient sequence $\{b_n\}$ is then determined by the following equation: \begin{equation}\label{b+K.b=0} \boldmath{b}+\boldmath{K b}=\boldmath{0}\,, \end{equation} where \begin{equation} K_{mn}:=\frac{1}{q_m}\sum_{k=1}^{\infty}(\chi_m,\phi_k)\,p_k \left\{ \begin{array}{c} \tanh\\ \coth \end{array} \right\} (p_k a)\,(\phi_k,\phi_n)\,. \end{equation} The two approaches~(\ref{C.a=0}),~(\ref{b+K.b=0}) are, of course, equivalent, however, it may be useful to combine them in order to get a good idea about the numerical stability of the solution. For instance, in the situation of~\cite{ESTV} the approximants of~(\ref{C.a=0}) approach the limiting values from above, while those referring to~(\ref{b+K.b=0}) are increasing.} \end{rem} \subsubsection{The Asymmetric Case} Let us pass now to the case, when the widths of the ducts are nonequal, $\nu\not=1$. In view of the mirror symmetry, we shall consider right-halfplane part of $\Omega$ only with the Neumann and Dirichlet condition on the segment $[-d_2,d_1]$ of the $y$-axis. The asymmetric case differs from the previous one just in the choice of transverse basis: now we can take \begin{eqnarray} \chi_{n}(y) &:=& \chi_n(y;\alpha_0) \qquad\mbox{if}\quad x\geq a \nonumber\\ \phi_{n}(y) &:=& \chi_n(y;\alpha_1) \qquad\mbox{if}\quad 0\leq x<a, \end{eqnarray} where $\chi_n(\cdot,\alpha_s)$ are of the form~(\ref{eigenfunctions}) with the norms $N_n^s$ given by~(\ref{norms}). The corresponding eigenvalues $\nu_n^0, \nu_n^1$ are then determined by \begin{eqnarray*} -\alpha_0 &=& \ell\,(\cot\ell d_1+\cot\ell d_2) \qquad\mbox{if}\quad x\geq a\\ -\alpha_1 &=& \ell\,(\cot\ell d_1+\cot\ell d_2) \qquad\mbox{if}\quad 0\leq x<a\,. \end{eqnarray*} (see~(\ref{spc})) and the overlap integrals are \begin{eqnarray}\label{ASoverlap} (\chi_m,\phi_n) &=& \frac{N_m^0 N_n^1}{\nu_m^0-\nu_n^1}\ \Big(\sqrt{\nu_n^1}\,\sin\sqrt{\nu_m^0}d_1\,sin\sqrt{\nu_m^0}d_2 \,\sin\sqrt{\nu_n^1}D\nonumber\\ &&\qquad\quad\ -\sqrt{\nu_m^0}\,\sin\sqrt{\nu_n^1}d_1,\sin\sqrt{\nu_n^1}d_2 \,\sin\sqrt{\nu_m^0}D\Big)\,. \end{eqnarray} The rest of the argument does not change and one has again to solve the equation~(\ref{C.a=0}) (respectively,~(\ref{b+K.b=0})). By a straightforward modification of the above argument, one can also check that the coefficient sequences have a faster than powerlike decay and that the equation can be solved by a sequence of truncations. \begin{figure}[!tb] \begin{center} \includegraphics[width=\textwidth]{fig02.eps} \end{center} \caption{Bound state energies {\em vs.} the halfwidth $\tilde a$ in the symmetric case for $\tilde\alpha_0=10^5\ (\square)$, $50\ (\circ)$, $10\ (\ast)$, $5\ (\times)$, $2\ (+)$, $0.5\ (\bullet)$.}\label{figEigenvalues} \end{figure} \begin{figure}[!tb] \begin{center} \includegraphics[width=\textwidth]{s_fig03.eps} \end{center} \caption{Ground-state eigenfunctions in the symmetric case for $a/d=0.15$.}\label{figEigenfunctions} \end{figure} \begin{figure}[!tb] \begin{center} \includegraphics[width=\textwidth]{fig04.eps} \caption{Narrow-window asymptotic power and coefficient as functions of $\alpha_0$.}\label{fig24} \end{center} \end{figure} \begin{figure}[!tb] \begin{center} \includegraphics[width=100mm]{s_fig05.eps} \caption{The eigenfunction of the second excited state for $\nu=1$, $a/d=5$, $\alpha_1=0$.}\label{figExcitedEigenfunction} \end{center} \end{figure} \begin{figure}[!tb] \begin{center} \includegraphics[width=\textwidth]{fig06.eps} \caption{The maximum bending of nodal lines of third eigenfunctions in the symmetric case as a function of~$\alpha_0$ for a fixed window width.}\label{figNodalLines} \end{center} \end{figure} \begin{figure}[!tb] \begin{center} \includegraphics[width=100mm]{fig07.eps} \includegraphics[width=100mm]{fig08.eps} \caption{Right-left conductivity as a function of $k$, $d_2$ for $d=\pi$, $a=2\pi$, $\alpha_1=0$.}\label{figConductivity} \end{center} \end{figure} \begin{figure}[!tb] \begin{center} \includegraphics[width=\textwidth]{s_fig09.eps} \caption{Probability flow patterns for $\tilde k=1.745$ and $\alpha_1=0$ in the symmetric situation.}\label{figFlow} \end{center} \end{figure} \subsection{Scattering} As we said in the opening of this section, the scattering can be treated in an analogous way. The incident wave is supposed to be of the form $e^{-i k_r x} \chi_r(y;\alpha_0)$, \ie, to come from the right in the $r$-th transverse mode; we have introduced the effective momentum $k_r:=\sqrt{k^2-\nu_r^0}$. We denote by $r_{rn}, t_{rn}$, respectively, the corresponding reflection and transmission amplitudes to the $n$-th transverse mode. Due to the mirror symmetry, we can again separate the symmetric and antisymmetric situation w.r.t. $x=0$ and to write \begin{equation}\label{amplitudes} r_{rn}=\frac{1}{2}(\rho_{rn}^s+\rho_{rn}^a)\,,\qquad t_{rn}=\frac{1}{2}(\rho_{rn}^s-\rho_{rn}^a)\,, \end{equation} where $\rho_{rn}^{\sigma}$, $\sigma=s,a$, are the reflection amplitudes in a half of our waveguide with the Neumann and Dirichlet condition at $x=0$, respectively. We use the following Ansatz \begin{equation} \begin{array}{rll} \Psi_{s/a}(x,y)&=\sum\limits_{n=1}^{\infty} \left(\delta_{rn}e^{-i k_n (x-a)} +\rho_{rn}^{s/a}e^{i k_n (x-a)}\right)\chi_n(y) &\quad\mbox{for}\quad x\geq a\\ &&\\ \Psi_{s/a}(x,y)&=\sum\limits_{n=1}^{\infty}a_n^{s/a} \left\{ \begin{array}{c} \frac{\cos p_n x}{\cos p_n a}\\ \\ \frac{\sin p_n x}{\sin p_n a} \end{array} \right\} \phi_n(y) &\quad\mbox{for}\quad 0\leq x<a \end{array} \end{equation} for the total energy $k^2$ of the incident wave in the $r$-th mode. The quantities $p_n:=\sqrt{k^2-\nu_n^1}$ are effective momenta in the ``interaction" region. Matching these functions smoothly at $x=a$ we arrive in the same way as above at the equation \begin{equation}\label{C.a=f} \boldmath{C a}=\boldmath{f}\,, \end{equation} where \begin{eqnarray} C_{mn} &:=& \left(i k_m+p_n \left\{ \begin{array}{c} \tan\\ -\cot \end{array} \right\} (p_n a)\right)(\chi_m,\phi_n)\\ f_m &:=& 2 i k_m \delta_{rm}, \end{eqnarray} where the index $r$ corresponds to the incident wave and the overlap integrals are given again by~(\ref{ASoverlap}). The reflection amplitudes are then given by \begin{equation} \rho_{rm}=-\delta_{rm}+\sum_{n=1}^{\infty} a_n (\chi_m,\phi_n)\,; \end{equation} they determine the full S-matrix via~(\ref{amplitudes}). A quantity of direct physical interest is rather the conductivity given by the Landauer formula. If we express it in the standard units $2 e^2/h$, it equals \begin{equation} G(k)=\sum_{m,n=1}^{[k]}\frac{k_n}{k_m}|t_{mn}(k)|^2\,, \end{equation} where $t_{mn}(k)$ are the coefficients~(\ref{amplitudes}). The summation runs over all open channels. Another physically interesting quantity is the probability flow distribution associated with the generalized eigenvector $\Psi=\Psi_s+\Psi_a$, which is defined in the standard way, \begin{equation} \vec j(\vec x)=2\im\left(\bar\Psi(\vec x)\nabla\Psi(\vec x)\right). \end{equation} \subsection{Numerical results} Since the spectrum behaves naturally at scaling transformations it is reasonable to solve the equations~(\ref{C.a=0}) and~(\ref{C.a=f}) in the natural non-dimensional quantities. We mark them by tilde and use them to label the axis in the figures, \eg, $a=d\tilde a$, $\alpha_s=\tilde\alpha_s/d$, $E=(\pi/d)^2\tilde E$ or $k=(\pi/d)\tilde k$. \subsubsection{Bound states} \paragraph{Eigenvalues:} Figure~\ref{figEigenvalues} shows the bound-state energies as functions of the ``window'' halfwidth $a$ for an ``empty window", $\alpha_1=0$. Several curves referring to different values of the barrier coupling constant $\alpha_0$ are plotted. In accordance with the general results of Section~\ref{Sec.existence} the energies decrease monotonously with the increasing ``window'' width and one can sandwich them between the estimates~(\ref{estE}). We also see that for a fixed $a$ the energies increase with respect to $\alpha_0$ and $\nu$; recall that their number increases as a function of $\alpha_0$ but it decreases as the waveguide becomes more asymmetric --- these facts are clear from~(\ref{estE}),~(\ref{estg}), and Lemma~\ref{approxbound}. It is illustrative to confront our results for large $\alpha_0$ with the energies computed in~\cite{ESTV} for the case which corresponds formally to $\alpha_0=\infty$. Comparing Figure~\ref{figEigenvalues} with Fig.~2 of the mentioned paper we see that our result for $\tilde\alpha_0=10^5$ is practically identical with the latter. \paragraph{Eigenfunctions:} The evolution of the ground-state wavefunction with respect to $\alpha_0$ for an empty window of the fixed halfwidth $\tilde a=0.15$ is illustrated on Figure~\ref{figEigenfunctions}. If the barrier tunneling is negligible, $\tilde\alpha_0=10^5$, the picture is indistinguishable from Fig.~3 in~\cite{ESTV}. As $\alpha_0$ becomes smaller we see how the wavefunction part penetrating the barrier grows. \paragraph{Threshold behaviour:} Consider again the empty--window case, $\alpha_1=0$. As a consequence of Theorem~\ref{WeakThm} we get for the ground-state energy for any fixed $\alpha_0$ and a narrow window the asymptotic formula (\cf~(\ref{expansion})) \begin{equation}\label{tresh} E(a)=\nu_{1}^0-c\,a^2+{\cal O}(a^3)\ ,\ \qquad c:=a^2 \alpha_{0}^2\ |\chi_{1}(0;\alpha_0)|^4. \end{equation} On the other hand, in the case of a window in the Dirichlet barrier, $\alpha_0=\infty$, it was conjectured in~\cite{ESTV} that we may suppose \begin{equation}\label{windowResult} E(a)=\left(\frac{\pi}{d}\right)^2-C(\nu)\,a^4+{\cal O}(a^5) \end{equation} as $a\to 0+$. The conjecture is supported by a two-sided asymptotic estimate~\cite{EV2}: there are positive $c_{1},c_{2}$ such that \begin{displaymath} -c_{1} a^4\leq E(a)-\left(\frac{\pi}{d}\right)^2\leq-c_{2} a^4 \end{displaymath} (for a generalization of this result to a larger number of windows and higher dimensions see~\cite{EV3}). Quite recently, a proof of (\ref{windowResult}) has been proposed by Popov~\cite{Pop}. This seems to be a paradox. In order to make sense of these considerations, we suppose that $E(a)=\nu_{1}^0-c a^{\beta}$ for small $a$ of an interval $0.016<\tilde a<\tilde a_{max}$ ($\tilde a_{max}=\tilde a_{max}(\alpha_0)$ is chosen in such a way to include the best correlated points), and investigate numerically the dependence of the coefficients $\beta$ and $c$ ($\tilde c=d^{\beta+2}c/\pi^2$) on $\alpha_0$. The powerlike asymptotic behaviour is confirmed when we redraw the first eigenvalue curves of Figure~\ref{figEigenvalues} in the logarithmic scale. The obtained dependence of the coefficients on $\alpha_0$ in the symmetric case $\nu=1$, is illustrated on Figure~\ref{fig24}. We see that the power reaches the values $\beta=2,4$ for small and large $\alpha_0$, respectively (a slight shift in the first graph is due the truncation; the convergence becomes very slow for small $a$). At the same time, the numerically found $c$ for small $\alpha_0$ coincides with that of~(\ref{tresh}). Of course, the asymptotical behaviour is governed by~(\ref{tresh}) for any finite $\alpha_0$. The above result says only that the transition from biquadratic to quadratic asymptotics occurs for large $\alpha_0$ at values of $a$ still smaller than those we have used. \paragraph{Nodal Lines:} In Figure~\ref{figExcitedEigenfunction} we plot the third eigenfunction. Its nodal lines are almost straight showing thus that the ``spikes'' at the window edges act almost as a hard barrier. On the other hand, a simple argument based on the reflection principle shows that the nodal lines cannot be straight. Closer inspection shows that they have the form of a bow bent outward. The maximum bending is shown on Figure~\ref{figNodalLines}. It decreases rapidly with the window width which confirms the tunneling nature of the effect. Nodal lines of higher eigenfunctions exhibit (as functions of~$\alpha_0$) irregularities connected with the changes in the number of the modes. \subsubsection{Scattering} \paragraph{Conductivity:} Figure~\ref{figConductivity} illustrates the evolution of the conductivity for the particle coming from the right and leaving to the left as a function of the momentum $k$ and the width $d_2$. We see that the perturbation, $-\alpha_0$ in the window, deforms the ideal steplike shape with jumps at transverse thresholds; deep resonances are clearly visible. For an almost impenetrable barrier, $\tilde\alpha_0=10^5$, we practically reproduce Fig.~5a of~\cite{ESTV}. \paragraph{Probability flow:} Examples of the quantum probability flow are shown on Figure~\ref{figFlow}. The flow patterns change with the momentum of the incident particle and the value of $\alpha_0$. They exhibit conspicuous vortices at the resonance energies which correspond to the ``trapped part'' of the wavefunction. An interesting phenomenon is illustrated on the first two graphs of Figure~\ref{figFlow}: for $\tilde\alpha_0=10^5$ there is a double vortex (corresponding to the sharp stopping resonance of Figure~\ref{figConductivity}), right-handed in the upper duct, while for $\tilde\alpha_0=50$ we get a left-handed vortex. The conductivity is small in these situations so the waveguide system is effectively closed for the particle transport. As $\alpha_0$ decreases the conductivity grows and the waveguide opens --- \cf~Figure~\ref{figFlow} for $\tilde\alpha_0=10,2$. \setcounter{equation}{0} \section{An Aharonov-Bohm cylinder} In the closing section we want to show now how the preceding considerations modify for a different geometry: we consider a nonrelativistic quantum particle living on the surface of an infinite straight cylinder of a radius $R$, which is threaded by a homogeneous magnetic field $\vec B$ parallel to the cylinder axis. We assume that the motion is further restricted by a $\delta$-barrier supported by a line parallel to the axis. \begin{figure}[!ht] \begin{center} \includegraphics[width=\textwidth]{fig10.eps} \caption{Cylindrical strip with a $\delta$ barrier in axial magnetic field.}\label{Cschm} \end{center} \end{figure} \subsection{General considerations} The configuration space is sketched on Figure~\ref{Cschm} where we indicate also how the coordinate system is chosen. In these coordinates we have \begin{equation} \tilde\Omega:=\{(x,y,z)\in\Real^3 |\ y^2+z^2=R^2\}\,. \end{equation} Choosing the gauge so that the electromagnetic potentials fulfill $\varphi(\vec x)\equiv 0$ and $\vec A(\vec x)=\frac{1}{2}\vec B\times\vec x$, the Hamiltonian acquires the form \begin{equation}\label{CHamiltonian} \tilde H_{\alpha}:=(-i\nabla+\vec A)^2 =-\Delta-i\frac{B}{2}(y\partial_z-z\partial_y)+\frac{B^2 R^2}{4} \end{equation} away of the barrier, where $B:=|\vec B|$; as before we put $\hbar=2m=1$, and also $e=-1$ having in mind an electron. The subscript $\alpha$ indicates the real function which defines the shape of the barrier as in the strip waveguide situation; it will enter the boundary condition (\ref{newCbc}) below. The vector potential has obviously the the angular component $A_{\varphi}\equiv A$ and it equals \begin{equation}\label{flux} A=\frac{1}{2}B R =\frac{\phi}{2\pi R}, \end{equation} where $\phi$ is the magnetic flux through the cylinder. Recall that in the rational units we use here, the natural unit of the magnetic flux is $\phi_0:= (2\pi)^{-1}$. Since we deal with a quantum system living on a surface it is natural to ``unfold" it and to study~(\ref{CHamiltonian}) on a planar strip with appropriate boundary conditions, namely \begin{equation}\label{newCbc} \psi(u,0+)=\psi(u,2\pi R-)=:\psi(u,0), \quad \psi_v(u,0+)-\psi_v(u,2\pi R-)=\alpha(u)\psi(u,0), \end{equation} where the subscript denotes again a partial derivative. To this aim, we introduce the unitary transformation $U:\sii(\tilde\Omega)\to\sii(\Real\times[0,2\pi R],du dv)$ by \begin{equation} (U\psi)(u,v):=\psi(u,R\cos\frac{v}{R},R\sin\frac{v}{R})\,, \end{equation} which maps $\tilde\Omega$ onto the strip $\Omega:=\Real\times[0,2\pi R]$; the operator $\tilde H_\alpha$ is then unitarily equivalent to \begin{equation}\label{newCHamiltonian} H_{\alpha}:=U\tilde H_{\alpha} U^{-1} =-\partial_u^2+(-i\partial_v+A)^2\,, \end{equation} with the domain \begin{equation}\label{CDomain} D(H_{\alpha}):=\left\{ \psi\in\sobii(\Omega)\,|\ \forall u\in\Real:\; {\rm b.c.\;} (\ref{newCbc})\; {\rm are\; satisfied}\, \right\}. \end{equation} We will need also the quadratic form $q_{\alpha}$ associated with $H_{\alpha}$. Its domain is $D(q_{\alpha}):=\sobi(\Omega)$ and \begin{eqnarray} q_{\alpha}[\psi] &:=& \int_{\Omega}|\nabla\psi|^2(u,v) du dv +\int_{\Real}\alpha(u)|\psi(u,0)|^2 du\nonumber\\ && - 2i A\int_{\Omega}(\bar\psi\partial_{v}\psi)(u,v) du dv +A^2\int_{\Omega}|\psi|^2(u,v) du dv\,. \end{eqnarray} As a comparison operator we employ again the one with $\alpha(u)\!=\!\alpha\!=\!const$ when we can solve the Schr\"odinger equation by separation of variables. We denote by $\{\nu_n\}_{n=1}^{\infty}$ and $\{\chi_n\}_{n=1}^{\infty}$ the (properly ordered) sequences of the transverse eigenvalues and the corresponding eigenfunctions, respectively. Since our system is now more complicated due to the presence of the magnetic field, we have to distinguish several possibilities: \\ [2mm] {\bf 1.}\quad \underline{No barrier}, $\alpha=0$: \begin{equation}\label{CEigenfunctions1} \forall\ell\in\Int:\qquad\tilde\chi_{\ell}(v)=\frac{1}{\sqrt{2\pi R}} \ e^{i\frac{\ell}{R}v}\,. \end{equation} The tilde marks the eigenfunctions corresponding to the eigenvalues $(\frac{\ell}{R}+A)^2$; to get $\{\nu_n\}$ one has to arrange thee latter into a ascending sequence. The respective eigenfunctions will be then denoted as $\{\chi_n\}$.\\ [2mm] {\bf 2.}\quad $\alpha\not=0$ \quad and \quad $2 R A\not\in\Nat$: \begin{equation}\label{CEigenfunctions} \forall n\in\Nat\setminus\{0\}:\qquad \chi_n(v)=N_n\, e^{-i A v}\left(e^{i\sqrt{\nu_n}v} -\frac{e^{i 2\pi R A}-e^{i 2\pi R\sqrt{\nu_n}}} {e^{i 2\pi R A}-e^{-i 2\pi R\sqrt{\nu_n}}}\, e^{-i\sqrt{\nu_n}v}\right)\,, \end{equation} where $N_n$ denotes the normalization factor chosen to make $\chi_n$ a unit vector in $\sii(0,2\pi R)$, \begin{equation} \begin{array}{c} |N_n|^2:= \frac{\sqrt{\nu_n}\ [1-\cos 2\pi R (\sqrt{\nu_n}+A)]} {4\pi R\sqrt{\nu_n}(1-\cos 2\pi R A\cos 2\pi R\sqrt{\nu_n}) +2\sin 2\pi R\sqrt{\nu_n}(\cos 2\pi R A-\cos 2\pi R\sqrt{\nu_n})}, \end{array} \end{equation} and the increasing sequence $\{\nu_n\}$ arises from the spectral condition \begin{equation}\label{Cspc} -\alpha=2\ell\left(\cot 2\pi R\ell -\frac{\cos 2\pi R A}{\sin 2\pi R\ell}\right). \end{equation} In analogy with Lemma~\ref{approxbound}(a) we have $\sqrt{\nu_n}\in\frac{1}{2 R}(n\!-\!1,n)$ for any $n\in\Nat\setminus\{0\}$. \\[2mm] {\bf 3.}\quad $\alpha\not=0$ \quad and \underline{integer flux}, $2 R A\in 2\Nat$: \\ The transverse eigenfunctions are of the form~(\ref{CEigenfunctions}), while the spectral condition~(\ref{Cspc}) changes to \begin{equation}\label{Cspceven} \alpha=2\ell\tan\pi R\ell\,. \end{equation} Moreover, for $\ell\in\frac{1}{R}\Nat$ we always get the trivial solutions \begin{equation}\label{CEigenfunctionstriv} \chi_n^{triv}(v)=\frac{1}{\sqrt{\pi R}}\,e^{-i A v} \sin \ell v\,, \end{equation} which are independent of $\alpha$. The roots $\{\tilde\nu_n\}$ of~(\ref{Cspceven}) satisfy the estimates $\sqrt{\tilde\nu_1}\in\frac{1}{2 R}(0,1)$ and $\sqrt{\tilde\nu_n}\in\frac{1}{2 R}(2 n\!-\!3,2 n\!-\!1)$ for $n\in\Nat\setminus\{0,1\}$. \\ [2mm] {\bf 4.}\quad $\alpha\not=0$ \quad and \underline{half-integer flux}, $2 R A\in 2\Nat+1$: \\ As in the two preceding cases the transverse eigenfunctions are still~(\ref{CEigenfunctions}) but the spectral condition~(\ref{Cspc}) changes now to \begin{equation}\label{Cspcodd} -\alpha=2 \ell\cot\pi R \ell\,. \end{equation} The wave functions of the trivial solutions, $\ell\in\frac{1}{2 R}(2\Nat\!+\! 1)$ are~(\ref{CEigenfunctionstriv}) again and the nontrivial roots of~(\ref{Cspcodd}) satisfy $\sqrt{\tilde\nu_n}\in\frac{1}{2 R}(2 n\!-\!2,2 n)$ for all $n\in\Nat\setminus\{0\}$. \\ \begin{rem}\label{Crems1}{\rm The integer and half-integer values here refer to the natural flux unit mentioned above. The essential instrument for proving the existence of bound states is the requirement $\chi_1(0)\not=0$ (compare, \eg, to Eq.~(\ref{Q=}) above). In the absence of a barrier, $\alpha=0$, the eigenfunctions are always positive (see~(\ref{CEigenfunctions1})) and $\chi_1(0)\not=0$ holds for $\alpha\leq\alpha_m:=-\frac{2\sin^2\pi R A}{\pi R}$. In the case of a (half-)integer flux we have to exclude the trivial solutions~(\ref{CEigenfunctionstriv}). It is an analogy of the trivial-part exclusion described in~Remark~\ref{rem1}; the difference is that the triviality now does not come from the waveguide geometry, but rather from the magnetic field, \ie, an external parameter. It is also clear that the ground-state eigenfunction of the class (\ref{CEigenfunctions}) can vanish at the barrier only if $\alpha>0$ and the flux is half-integer.} \end{rem} In analogy with Lemma~\ref{approxbound} we have \begin{lemma}\label{Capproxbound} Suppose that $2RA\not\in 2\Nat+1$ or $\alpha\leq0$. Then the function $\alpha\mapsto\nu_1(\alpha)$ is strictly increasing and continuous. \end{lemma} The ``unperturbed" Green's function is the same as in the case of double waveguide (\cf~(\ref{green0})); one has only to substitute the present transverse eigenfunctions and eigenvalues. After this preliminary we can easily derive sufficient conditions under which a local perturbation of the barrier coupling parameter induces existence of bound states. The argument mimics that of Theorem~\ref{existence}, the only difference being an additional requirement of the magnetic field. \begin{thm}\label{Cexistence} Assume {\em (i)} $\;\alpha\!-\!\alpha_0\in\si_{loc}(\Real)$, \\[1mm] {\em (ii)}$\;\alpha(u)\!-\!\alpha_0={\cal O}(|u|^{-1-\varepsilon})$ for some $\varepsilon>0$ as $|u|\to\infty$,\\[1mm] {\em (iii)} the flux is not half-integer, $2RA\not\in 2\Nat+1$, if $\alpha_0>0$. \\[1mm] Then $H_{\alpha}$ has at least one isolated eigenvalue below its essential spectrum, $\sigma_{ess}(H_{\alpha})= [\nu_1(\alpha_0),\infty)$, provided $\int_{\Real}(\alpha(u)\!-\!\alpha_0)du<0$. \end{thm} \begin{figure}[!htb] \begin{center} \includegraphics[width=\textwidth]{s_fig11.eps} \end{center} \caption{The quantum probability flow on the cylinder surface for $\tilde A=0.5$, $\tilde k=1.705$ and $\tilde\alpha_1=10^{-5}$.}\label{CfigFlow} \end{figure} \subsection{An example} We shall illustrate the above consideration on a ``rectangular-well" example analogous to that of Section~\ref{Sec.Example}. Most of the argument proceeds as there, one has just to use the different eigenfunctions and to recompute for them the overlap integrals: \begin{eqnarray}\label{Coverlap1} (\chi_m,\phi_n) &=& \frac{8\,\bar N_m^0 N_n^1} {(e^{-i 2\pi R A}-e^{i 2\pi R\sqrt{\nu_m^0}}\ ) (e^{i 2\pi R A}-e^{-i 2\pi R\sqrt{\nu_n^1}}\ )\, (\nu_m^0-\nu_n^1)}\cdot\nonumber\\ &&\cdot\Bigl[\sqrt{\nu_m^0}\sin2\pi R\sqrt{\nu_n^1} \ (\cos 2\pi R A-\cos 2\pi R\sqrt{\nu_m^0}\ )\nonumber\\ &&\ -\sqrt{\nu_n^1}\sin 2\pi R\sqrt{\nu_m^0} \ (\cos 2\pi R A-\cos 2\pi R\sqrt{\nu_n^1}\ )\Bigr] \end{eqnarray} for $\alpha_0\not=0\not=\alpha_1$ with $m,n\in\Nat\setminus\{0\}$, and \begin{equation}\label{Coverlap2} (\chi_m,\phi_{\ell})=\frac{4\,\bar N_m^0\,\sqrt{\nu_m^0}} {\sqrt{2\pi R}\,\left(\nu_m^0-(\frac{\ell}{R}+A)^2\right)}\ \left(\sin 2\pi R A+i\cos 2\pi R\sqrt{\nu_m^0}\,\right) \end{equation} for $\alpha_0\not=0$, $\alpha_1=0$ with $m\in\Nat\setminus\{0\}$, $\ell\in\Int$. Note that in the latter case one has to substitute the {\em ordered} basis $\{\phi_n\}_{n=1}^{\infty}$ (together with the corresponding eigenvalues) into Ansatz~(\ref{Ansatz}) to make the numerical procedure of cut-off approximations convergent. Unless $\alpha_1<0$ we have to exclude here the possibility $2 R A\in 2\Nat\!+\!1$ again (\cf, \eg,~(\ref{estg})). Next we can restrict our attention only to the situation when $2 R A\in [0,1)\cup(1,2)$ because $A$ appears in the overlap integrals~(\ref{Coverlap1}), (\ref{Coverlap2}) and in the spectral condition~(\ref{Cspc}) as an argument of the periodic functions $\sin$, $\cos$; the integrals and the transverse eigenvalues are the only quantities which affect the equations~(\ref{C.a=0}) and~(\ref{C.a=f}). In fact, we can take $2 R A$ from $[0,1)$ only because the replacement $2 R A\mapsto 2-2 R A$ in~(\ref{Coverlap1}) and~(\ref{Coverlap2}) (the spectral condition~(\ref{Cspc}) does not change at all) is equivalent to the exchange $A\mapsto -A$ which coincides with the conjugation of Hamiltonian~(\ref{newCHamiltonian}). It is well known fact that such an operator has the same energies while the corresponding eigenfunctions are given by a simple conjugation. As an example, the evolution of the quantum probability flow w.r.t. $\alpha_0$ is illustrated on Figure~\ref{CfigFlow}.
\section{introduction} Recently, there has been growing interest about the possibility of the formation of a disoriented chiral condensate (D$\chi$C) in heavy ion collisions \cite{And,KK,An1,An2,BK,Bj1,Bj2,RW1,KT,RW2,BKT1,BKT2,GGP,GM,CBNJ,CKMP,AHW,R,C,Bj3,old}. In an ultrarelativistic heavy ion collision, some region may thermalize at a temperature high enough so that chiral symmetry is restored in the region. As the system cools sufficiently rapidly back through the transition temperature, the chiral restored state is unstable as small fluctuations in any chiral direction ($\sigma,\vec{\pi}$) will grow exponentially. This can create regions where the pion field has a macroscopic occupation number. It should be stressed that this scenario is not derivable directly from the underlying theory of QCD and contains a number of untested dynamical assumptions, principally that the cooling is rapid. While the failure of the system to form a D$\chi$C cannot be used to rule out that the system has reached the chiral restoration temperature (as the scenario described above is not derivable directly from QCD), observation of D$\chi$C formation would be a clear signal for chiral restoration at high temperature. There are clear signatures of D$\chi$C formation provided a single large domain is formed, containing a large number of pions. For example, one expects an excess in low $p_T$ pion production as the characteristic momentum of a pion from a large region is small. Such a signal works even if multiple regions of D$\chi$C are formed provided each region is large, but it is not decisive since one could imagine some other collective low energy effects which produce low $p_T$ pions. On the other hand, since the pions formed in a D$\chi$C, being essentially classical, form a coherent state, this coherent state has some orientation in isospace, and all of the pions in the domain are essentially maximally aligned (given the constraints of quantum mechanics) and point in the same isospin direction. If there are a large number of pions in the domain, this implies a distinctive distribution of $R$, the ratio of neutral to total pions in the domain \cite{And,KK,An1,An2,Bj1,Bj2,RW1,R}: \begin{equation} f_0(R) \, = \, \frac{1}{2 \sqrt{R}}. \label{PofR} \end{equation} In contrast, the distribution from uncorrelated emissions is narrowly peaked about $1/3$ with a variance, $\langle R^2 \rangle - \langle R \rangle^2 = \frac{2}{9{\cal N}}$ where $\cal N$ is the total number of pions, and approaches a delta function at $R=1/3$ when ${\cal N}\to\infty$. Since these two distributions are dramatically different, this provides a clear signature for single domain D$\chi$C formation, provided one can kinematically separate the pions from the D$\chi$C from other pions in the system. Unfortunately, this signature depends critically on the assumption that a single large domain of D$\chi$C is formed, which is {\it a priori} rather unlikely. One expects the domains are of characteristic size $\tau$, the exponential growth time of the pion fluctuation in the unstable chiral restored state\cite{GGP,GM}. Since the size of the fireball is much larger than typical QCD length scales, it seems unlikely that $\tau$ would be of the size of the fireball, and thus formation of a single D$\chi$C domain is improbable. Formation of multiple domains of D$\chi$C, however, tends to wash out the $R$ distribution described by Eq.~(\ref{PofR}). As the pions emerging from different domains cannot be distinguished kinematically, by the central limit theorem the $R$ distribution will approach a normal distribution peaked at $R=1/3$. This normal distribution may be distinguished from the normal distribution arising from uncorrelated emission; the case of multiple domains of D$\chi$C will have a substantially larger variance. Unfortunately, there is an important practical limitation which makes it difficult to exploit the $R$ distribution as a signature. Even under the most optimistic of scenarios, the total number of pions coming from D$\chi$C's will be a small fraction of the total number of pions. If one includes all pions produced in the reaction, the signal due to the pions from the D$\chi$C will presumably be overwhelmed. One can apply a low $p_T$ cut to suppress the noise due to incoherently emitted pions. However, even with the low $p_T$ cut, the noise may still be severe, as both the signal and the noise peak at $R=1/3$. In this paper, we suggest cuts which may dramatically enhance the signal-to-noise ratio. We study the conditional probability distribution of $R$ given only for the events in which the $k$ pions with the lowest $p_T$ are all neutral, and we will show that the expectation value of $R$ is shifted away from $1/3$. Since incoherent emission will result in a very narrow peak around $R=1/3$, any such shifts should be easily observable. Moreover, one can make successive cuts by increasing the value of $k$, and enhance the signal in each successive step. This paper is organized as follows. We will start with the simple scenario and study single domain D$\chi$C formation in Sec.~II. In Sec.~III the effects of multi-domain formation and the noises due to incoherently emitted pions will be studied, while potential experimental application and limitations will be discussed in Sec.~IV. \section{domain with non-zero isospin} We are going to start with an unrealistic simple scenario and make it more realistic later on in our discussion. We will consider a single domain which is described by an isosinglet density matrix. (In general, a D$\chi$C is not a pure state; the wavefunction of the pion coherent state at the core of the ``fireball'' of the heavy ion collision is entangled with the energetic emission at the edge of the ``fireball''.) Such a state can be written as \begin{equation} \rho = \sum_{n,I,I_z} c_{nI}{(-1)^{I_z}\over(2I+1)}\, |n,I,I_z\rangle \langle n,I,I_z|, \label{domain} \end{equation} with $\sum_{n,I} c_{nI} = 1$. Note that the coefficients $c_{nI}$ are real, positive and do not depend on $I_z$ (as the full state is assumed to be isoscalar). The probability distribution of such a mixed state in the isospace is \footnote {Here we are assuming that the typical $I$ is much less than $\langle n \rangle$.} \begin{equation} d^2P(\theta,\phi) = \sum_{n,I} c_{nI}/(2I+1) \, \sum_{I_z} |Y_{II_z}(\theta,\phi)|^2 \; \sin\theta d\theta\, d\phi = \sin\theta d\theta\, d\phi / 4\pi, \qquad dP(\theta) = \sin\theta d\theta/2, \end{equation} where the angles $(\theta,\phi)$ are defined such that a unit vector in isospace is $(r_x,r_y,r_0) = (\sin\theta \cos\phi, \sin\theta \sin\phi, \cos\theta)$. The probability distribution is uniform, as the condensate is equally likely to point at any direction on the two-dimensional sphere, as demanded by isospin symmetry. The number operator of neutral pions in the condensate is given by \begin{equation} n_0 = a_0^\dag a_0 = \vec a^\dag \cdot \vec a \cos^2\theta = \langle n \rangle \cos^2 \theta, \end{equation} with $\vec a = (a_x,a_y,a_0)$ is a vector of hermitian annihilation operators which annihilate $\pi_x=(\pi_+ +\pi_-)/\sqrt{2}$, $\pi_y=(\pi_+ - \pi_-)/\sqrt{2} i$, and $\pi_0$, respectively, and $\langle n \rangle = \sum_{n,I} c_{nI} n$ is the expected number of pions. Since $R=n_0/\langle n \rangle$ \footnote{ We are assuming that $\langle n \rangle \gg 1$.}, one can easily calculate the probability distribution of $R$. \begin{equation} dP = \textstyle{1\over2}\sin\theta d\theta = \textstyle{1\over2} d\cos\theta = d\sqrt{R} = \textstyle{1\over2} R^{-1/2} dR. \end{equation} By defining $dP \equiv f_0(R) \, dR$ (the subscript ``0'' stands for $I=0$), one has, in the limit that $\langle n \rangle$ is large, \begin{equation} f_0(R) = \textstyle{1\over2} R^{-1/2}, \end{equation} recovering Eq.~(\ref{PofR}). The distribution is plotted in Fig.~1a. It is obvious that the shape is qualitatively different from the Poisson--Gaussian distribution due to incoherent emissions. The expectation value of $R$ is $1/3$, \begin{equation} \langle R \rangle_0 \equiv \int_0^1 R f_0(R) dR = 1/3, \label{1/3} \end{equation} which has the simple interpretation that it is equally likely for the pion to be a $\pi_0$, $\pi_+$ or $\pi_-$, and hence on average a third of the pions are neutral. This distribution is a consequence of the fact that we have assumed the D$\chi$C to be an isosinglet, a reasonable assumption on physical grounds. However, let's consider the distribution of $R$ after the D$\chi$C emits a single $\pi_0$. The density matrix after the emission, which can be written as $\lambda \; a_0 \, \rho \, a_0^\dag$, where $\lambda$ is a normalization constant, $\rho$ is the density matrix defined in Eq.~(\ref{domain}) and $a_0$ annihilates a neutral pion. This new density matrix is not an isosinglet. It is straightforward to show that the probability distribution for this state is \begin{equation} dP = {\textstyle{1\over2}\sin\theta \cos^2\theta d\theta\over \int\textstyle{1\over2}\sin\theta \cos^2\theta d\theta} = {\textstyle{1\over2} R \cdot R^{-1/2} dR \over \int_0^1 \textstyle{1\over2} R \cdot R^{-1/2} dR} \end{equation} or equivalently, \begin{equation} f_1(R) \equiv f(R|\hbox{1st pion is neutral}) = R f_0(R) \bigg/\int_0^1 R f_0(R) dR = \textstyle{3\over2} R^{1/2}. \end{equation} The distribution $f_1(R)$ is plotted in Fig.~1b, which is drastically different from $f_0(R)$. The distribution is skewed towards the high end, while $f_0(R)$ is skewed towards the low end. Moreover, the expectation value of $R$ is clearly pushed up: \begin{equation} \langle R \rangle_1 \equiv \int_0^1 R f_1(R) dR = 3/5. \label{3/5} \end{equation} So we have arrived at the intriguing conclusion that, if the ``first pion'' emitted from a isosinglet D$\chi$C is neutral, 60\% of the pions subsequently emitted from the D$\chi$C are neutral, a huge enhancement from the original expectation of 33\%. \bigskip \begin{figure} \epsfig{file=pi0.eps, width=3.4in} \epsfig{file=pi1.eps, width=3.4in} \hskip 115pt Fig.~1a \hskip 220pt Fig.~1b \bigskip \epsfig{file=pi2.eps, width=3.4in} \epsfig{file=pi3.eps, width=3.4in} \hskip 115pt Fig.~1c \hskip 220pt Fig.~1d \bigskip \caption{The probability distribution functions $f_k(R)$ for different values of $k$. Fig.~1a, b, c and d are $f_0(R)$, $f_1(R)$, $f_2(R)$ and $f_3(R)$, respectively. } \end{figure} \bigskip This extraordinary statement certainly deserves more discussion. First, what is our criterion to decide which is the ``first pion''? The answer is simple: it can be any criterion. It does not matter as long as it is {\it a priori\/} equally likely to be a $\pi_0$, a $\pi_+$ or a $\pi_-$. The derivation just depends on our removing a neutral pion from the isosinglet D$\chi$C. It can be the first pion emitted in time, or the last one emitted in time, or even the 17th emitted in time. The criterion can also be unrelated to the order of emission. For example, we can choose the ``first pion'' to be the one with the smallest polar angle. One can use any of these criteria to identify the ``first pion'', and if it turns out to be neutral, then the $R$ distribution of subsequent emissions is always given by $f_1(R)$, provided we are in the large number limit. However, this is only true in this idealized scenario, when all the pions are coming from a simple D$\chi$C domain. In reality, some of the pions come from incoherent emission, and if the ``first pion'' turns out to be incoherently emitted, the expectation value of $R$ of the remaining pions is still going to be $1/3$, not $3/5$. As a result, we want to choose our criterion in such a way that the ``first pion'' is likely to originate from the D$\chi$C and not from incoherent emissions. Since D$\chi$C pions by hypothesis have low $p_T \sim 1/L$, where $L$ is the size of the domain, a natural choice is to use the pion with the lowest $p_T$ as our ``first pion''. After clarifying the meaning of the term ``first pion'', we move on to discuss the physical origin of the modification of the probability distribution of $R$. In a nutshell, we are seeing the physics of (iso)spin alignment due to Bose condensation. To illustrate the point, let us first consider the following apparently unrelated Stern--Gerlach experiment. Consider a large number of massive spin-1 particles, which for concreteness will be called deuterons. Initially they are all polarized along a randomly chosen direction $\vec n$, which is {\it a priori\/} equally likely to be any direction in three dimensional space. In other words, $\vec S \cdot \vec n = 0$ for all the deuterons. Now let us pick one of these deuterons and pass it through a Stern--Gerlach spectrometer which measures $S_z$, the spin along the $z$-axis. What is the probability that the measurement gives $S_z=0$? The answer is clearly $1/3$, as the cases for $S_z=+1$, 0 and $-1$ are equally likely. On the other hand, if the measurement on the first deuteron gives $S_z=0$, what is the conditional probability for the next deuteron to pass through the Stern--Gerlach spectrometer also to be measured to have $S_z=0$? The answer this time is no longer $1/3$. The spins of all the deuterons are aligned along the same direction $\vec n$, and that the first deuteron is measured to have $S_z=0$ suggests $\vec n$ is more probable to be more or less aligned along $\vec z$ than otherwise. As a result, the conditional probability is no longer $1/3$, but can be easily shown to be $3/5$, which is exactly the predicted value for $\langle R \rangle_1$ in Eq.~(\ref{3/5}). The situation for a single domain of D$\chi$C is analogous, with isospin aligned pions instead of spin aligned deuterons. By construction, the pions in a D$\chi$C domain are isospin aligned, and by the same analysis, we have shown that the knowledge of the ``first pion'' being neutral can dramatically modify the conditional probability distribution of $R$. One can also consider the conditional probability distribution of $R$ in the case that the ``first pion'' is charged. Note that \begin{equation} f_0(R) = \textstyle{1\over3}\big(f(R|\hbox{1st pion is a $\pi_+$}) + f(R|\hbox{1st pion is a $\pi_-$}) + f(R|\hbox{1st pion is a $\pi_0$})\big), \end{equation} and hence, since $f_1(R) = f(R|\hbox{1st pion is a $\pi_0$})$, \begin{equation} \tilde f(R) \equiv f(R|\hbox{1st pion is charged}) = \textstyle{3\over2} f_0(R) - \textstyle{1\over2} f_1(R) = \textstyle{3\over4} (1-R) R^{-1/2}. \end{equation} The expectation value of $R$, given that the ``first pion'' is charged, can be easily shown to be $1/5$. As a consistency check, one can calculate $\langle R \rangle$, the expectation value of $R$ regardless of the species of the ``first pion''. Since the ``first pion'' is twice as likely to be charged as to be neutral, \begin{equation} \langle R \rangle_0 = \textstyle{1\over3}(\textstyle{3\over5} +2 \times \textstyle{1\over5}) = \textstyle{1\over3}, \end{equation} agreeing with Eq.~(\ref{1/3}). Lastly, we will study the conditional probability distribution of $R$ given that the $k$ pions with the lowest $p_T$, which will be hereafter referred to as the ``first $k$ pions'', are all neutral. It is straightforward to show that in this case \begin{equation} dP = {\textstyle{1\over2}\sin\theta \cos^{2k}\theta d\theta\over \int\textstyle{1\over2}\sin\theta \cos^{2k}\theta d\theta} = {\textstyle{1\over2} R^k \cdot R^{-1/2} dR \over \int_0^1 \textstyle{1\over2} R^k \cdot R^{-1/2} dR}. \end{equation} and \begin{equation} f_k(R) \equiv f(R|\hbox{1st $k$ pions are all neutral}) = R^k f_0(R) \bigg/\int_0^1 R^k f_0(R) dR = (k+\textstyle{1\over2}) R^{k-1/2}. \end{equation} The distributions $f_2(R)$ and $f_3(R)$ are plotted in Fig.~1c and d, respectively. One can see that as $k$ increases, the distribution is more and more skewed towards the high end. As a result, the expectation value of $R$ increases with $k$. \begin{equation} \langle R \rangle_k \equiv \int_0^1 R f_k(R) dR = (2k+1)/(2k+3). \label{R} \end{equation} It is useful to define $Q$ as the ratio of the number of $\pi_+$ to the number of total pions emitted. By symmetry it is also the ratio of the number of $\pi_-$ to the number of total pions emitted, and since $R+2Q=1$, \begin{equation} \langle Q \rangle_k = 1/(2k+3). \label{Q} \end{equation} From the above analysis, the prescription to enhance the collective signal is quite clear. One should make successive cuts on the data sample on the condition that the $k$ pions with the lowest $p_T$ are all neutral, and measure $\langle R \rangle_k$ after each cut to see if it increases as predicted in Eq.~(\ref{R}). This result, however, depends on the assumption that we have only a single domain of D$\chi$C without any contamination due to incoherent pion emissions. Since this assumption is unrealistic for heavy ion collision experiments, the scenario we studied in this section is only an idealized situation. In the next section, we will discuss more realistic scenarios. \section{The effects of multi-domain formation and incoherent emissions} The scenario considered in the last section is highly unrealistic in at least two ways. First, as discussed in the introduction, single domain D$\chi$C formation is highly unlikely. For a realistic treatment one must study D$\chi$C formation with more than one domain, each pointing in a different direction in the isospace. Moreover, we have neglected the effect of incoherently emitted pions, which have very important effects. If the neutral ``first pion'' is incoherently emitted, the $R$ distribution of the remaining pions is described by $f_0(R)$, instead of $f_1(R)$ when the ``first pion'' comes from the D$\chi$C. In this section, we will incorporate these two effects and see how the predictions above are modified. We will study the expectation value of $R$, or equivalently the expectation of $Q$, for a situation described by the following parameters. The coherent fraction $\chi$ is the fraction of pions which originate from D$\chi$C domains, so that when $\chi=1$, all pions are coherently emitted, and when $\chi=0$, all pions are incoherently emitted. We will consider the case where there are $N$ domains, all containing an equal number of pions\footnote {This assumption of all domains having the same number of pions is unrealistic but is made for illustrative purposes. The effects of unequal domain sizes will be briefly discussed below.}, which will be assumed to be large. Each domain is described by an isosinglet density matrix, but the isospins of pions in different domains are uncorrelated. Now the question is: if the ``first $k$ pions'' in this channel are all neutral, what are the expectation values of $R$ and $Q$ among the rest of the pions? The answer turns out to be the following expression: \begin{mathletters} \begin{equation} \langle R \rangle = \textstyle{1\over3} + 2\Delta, \qquad \langle Q \rangle = \textstyle{1\over3} - \Delta. \end{equation} The shift $\Delta$ is given by \begin{equation} \Delta = \chi \sum_{j=0}^k P_j \, ({1\over 3} - {1\over 2j+3}) = \chi \big({1\over 3} - \sum_{j=0}^k P_j \, {1\over 2j+3} \big), \end{equation} where \begin{equation} P_j = {k \choose j} \; p^j \, (1-p)^{k-j}, \qquad p=\chi/N. \end{equation} \label{result} \end{mathletters} Each term in this formula has a simple interpretation: $\bullet$ The expectation value $\langle R \rangle$ is always $1/3$ for the incoherently emitted pions. Only the pions coming from the domains are affected by isospin alignment; hence the outstanding factor of $\chi$. $\bullet$ Each coherently emitted pion comes from one of the domains, which will be called domain X. How many of the ``first $k$ pions'' also come from domain X? The probability for each pion coming from domain X is $p=\chi/N$, and the probability that $j$ of the ``first $k$ pions'' coming from domain X is $P_j = {k \choose j} \, p^j \, (1-p)^{k-j}$. $\bullet$ Given that $j$ of the ``first $k$ pions'' is coming from domain X, the conditional expectation value of $Q$ decreases from $1/3$ to $1/(2j+3)$, while the conditional expectation value of $R$ increases by twice the above quantity. In passing, we note that $\Delta$ can also be expressed as an integral or the hypergeometric function $_2F_1$: \begin{eqnarray} \Delta &=& \chi \Big({1\over 3} - {1\over \sqrt{p^3}} \int_0^{\sqrt{p}} dz \; z^2 (z^2 + 1 - p)^k \Big) \nonumber \\ &=& \chi \Big({1\over 3} - {\cos^{2k+3}\Theta \over \sin^3\Theta} \int_0^\Theta d\vartheta\; {\sin^2\vartheta \over \cos^{2k+4}\vartheta} \Big), \qquad \tan^2\Theta = {\chi/N \over 1- \,\chi/N} \nonumber \\ &=& {1\over 3} \chi \Big(1-(1-p)^k \, _2F_1[{3\over2},-k,{5\over2},{-p\over 1-p}] \Big). \end{eqnarray} In Fig.~2, we have made contour plots of $\Delta=1/30$ (such that $\langle R \rangle = 0.4$ and $\langle Q \rangle = 0.3$), for $k=1,\dots,5$ in the $(\chi,1/N)$ parameter space. The horizontal axis is the coherent fraction $\chi$ while the vertical axis is $1/N$ where $N$ is the number of domains. Both $\chi$ and $1/N$ range from 0 to 1. Thus, for example, with 3 domains and $\chi=0.6$, in order to have $\Delta \geq 1/30$ we must have $k\geq 3$. Equations (\ref{result}) illustrate the main results of this paper. One can see that, without any D$\chi$C formation, $\chi=0$ (corresponding to the left edge of Fig.~2), $\Delta$ vanishes, and $\langle R \rangle = \langle Q \rangle = 1/3$ as expected. The bottom edge of the plot corresponds to $N\to\infty$ and also gives $\Delta = 0$ for any finite value of $k$. The shift $\Delta$ is largest for a single domain of D$\chi$C without any noise due to incoherently emitted pions, {\it i.e.}, when $\chi=N=1$ (the top right corner of the contour plots), giving $\Delta = 1/3 \, - \, 1/(2k+3)$ and reproducing Eqs.~(\ref{R}) and (\ref{Q}). For fixed values of $(\chi,N)$, $\Delta$ increases with $k$, accounting for the spreading of the parameter space with $\Delta > 0.4$ as $k$ increases from 1 to 5 in Fig.~2. \bigskip \begin{figure} \hskip 120pt $1/N$ \centerline{ \epsfig{file=newspread.eps, width=3.4in} $\chi$} \bigskip \caption{Contour plots of $\Delta = 1/30$ (such that $\langle R \rangle = 0.4$ and $\langle Q \rangle = 0.3$) for different values of $k$. The horizontal axis is the coherent fraction $\chi$, while the vertical axis is $1/N$ where $N$ is the number of domains. Both $\chi$ and $1/N$ range from 0 to 1. The curves, from top right to bottom left are for $k=1$, 2, 3 4 and 5, respectively. The shift $\Delta$ is larger than $0.4$ above the curves and smaller than $0.4$ below the curves. } \end{figure} \bigskip One expects that when the number of D$\chi$C domains is large ($N\gg1$) or when most of the pions are incoherently emitted ($\chi\ll 1$), it will be difficult to observe clear signals of D$\chi$C formation. However, in such situations $\chi/N$ is small and $\Delta$ is dominated by the $j=1$ term (the $j=0$ term always identically vanishes) and \begin{equation} \Delta = {2 \chi^2 k \over 15 N} + {\cal O}({\chi^3\over N^2}). \label{small} \end{equation} Thus a large $k$ may make up for a small coherent fraction $\chi$, or a large number of domains $N$, and enhance $\Delta$, which describes the shift of $\langle R \rangle$ and $\langle Q \rangle$ from $1/3$, to an experimentally measurable magnitude. From the form of Eq.~(\ref{small}), one expects this shift to be substantial whenever $k \sim N/\chi^2$. However, even for a value of $k$ as small as $N/4\chi^2$, $\Delta = 1/30 + {\cal O} (\chi^3/N^2)$, which translates to $\langle R \rangle = 0.4$ and $\langle Q \rangle = 0.3$ --- a substantial deviation from the incoherent case. This suggests one should make successive cuts for events where the $k$ pions with lowest $p_T$ are all neutral, and study $\langle R \rangle$ after each cut. An increase of $\langle R \rangle$ with $k$ would suggest that D$\chi$C domains are formed. Equation (\ref{small}) appears to suggest that one can increase $\Delta$ to an arbitrarily large magnitude by choosing a sufficiently large value of $k$. Of course this is not true. Equation (\ref{small}) is obtained as the leading term in a $\chi/N$ expansion, but when $k \to \infty$, this expansion breaks down as terms of higher order in $\chi/N$ are enhanced by factors of $k \choose j$. We can easily see that \begin{equation} \Delta \to \textstyle{1\over3}\chi, \quad \langle R \rangle \to \textstyle{1\over3}+\textstyle{2\over3}\chi, \quad \langle Q \rangle \to \textstyle{1\over3}(1-\chi), \qquad k\to\infty \hbox{ with $\chi$ and $N$ fixed.} \label{limit} \end{equation} In other words, our signal enhancement scheme is fundamentally limited by the amount of noise due to incoherently emitted pions. When $\chi$ is small, most of the pions are incoherently emitted, and for them, $\langle R \rangle$ is always around $1/3$ regardless of what cuts one makes. On the other hand, the large $k$ limit of $\langle R \rangle$ does not depend on $N$, the number of D$\chi$C domains. Recall that we have several distinctive signatures, like the $R$ distribution in Eq.~(\ref{PofR}) and the conditional expectation values for $R$ described in the previous section, for a single domain of D$\chi$C, where all the pions in the D$\chi$C are isospin aligned. With multi-domain formation, where the pions in different domains may point to different directions in isospace, the effect of isospin alignment is greatly washed out. However, given that the ``first $k$ pions'' are all neutral with $k\gg N$, it is probabilistically extremely likely that each of the $N$ domains is the origin of some of these ``first $k$ pions''. As a result, each of these $N$ domains are well-aligned along the $\pi_0$ direction, and hence also well-aligned with each other. As $k\to\infty$, the $N$ domains look more and more like a single big domain in the $\pi_0$ direction, which is the case where the signature is the most dramatic. In short, with large $k$, our cuts are picking out the events where the signals are the strongest, and hence resulting in a large signal-to-noise ratio. In above we have assumed the sizes of all $N$ domains are identical for illustrative purposes. A more realistic treatment would have $N$ domains, each with different sizes $p_i$, $1\leq i \leq N$, such that $\sum_i p_i = \chi$. (The size of a domain is defined to be the fraction of pions which originate from this particular domain.) Then Eqs.~(\ref{result}) are generalized to \begin{equation} \Delta = {1\over 3} \chi - \sum_{i=1}^N p_i \; \sum_{j=0}^k {k \choose j} p_i^k (1-p)^{k-j} \; {1\over 2j+3}. \end{equation} In the weak signal limit, {\it i.e.}, when all the $p_i$'s are small, Eq.~(\ref{small}) gets modified to \begin{equation} \Delta = {2k \over 15} \sum_{i=1}^N p_i^2 + {\cal O}({\chi^3\over N^2}) = {2k \over 15} \overline p + {\cal O}({\chi^3\over N^2}) , \end{equation} where $\overline p = \sum_i p_i^2$ has the following nice interpretation: $\overline p$ is the average over all pions (both coherently and incoherently emitted) of the sizes of the originating domains, which is $p_i$ for a pion from domain $i$ and zero for an incoherently emitted pion. For $N$ domains of equal sizes, $\overline p = \chi^2/N$ and Eq.~(\ref{small}) is recovered. Again we see that $\Delta$ grows linearly with $k$ in the weak signal limit. As $k\to\infty$ the shift $\Delta$ is again limited by the bound (\ref{limit}), which applies also for the cases of unequal domain sizes. \section{summary} To recapitulate, we suggest the following experimental procedures: $\bullet$ Count the number of neutral and charged pions {\it event by event\/} from heavy ion collision experiments and measure their individual transverse momenta and rapidities. $\bullet$ Apply a low $p_T$ cut to suppress the noise due to uncorrelated pion emission. $\bullet$ Bin the events in different rapidity windows. $\bullet$ In each rapidity window, calculate the expectation value $\langle R \rangle$. $\bullet$ Make a cut to retain only events where the pion with the lowest $p_T$ is neutral. $\bullet$ Calculate, in each rapidity window, the expectation value $\langle R \rangle$ for all remaining pions in all events which survive the cut. $\bullet$ Make another cut on the surviving events to retain only those where the pion with the second lowest $p_T$ is also neutral. $\bullet$ Again, calculate in each rapidity window the expectation value $\langle R \rangle$ for all remaining pions in all events which survive the cuts. $\bullet$ Repeat the above prescription of making successive cuts to retain only events in which the pion with the next lowest $p_T$ is also neutral, and calculate $\langle R \rangle$ for each rapidity window after each cut. If we find $\langle R \rangle$ deviates from $1/3$ then we are seeing signatures from D$\chi$Cs. Note that this prescription requires reconstructions of $p_T$'s of individual pions, both charged and neutral. We have also presumed that the coherent fraction $\chi$ and the number of domains formed $N$ are roughly the same for each event. (More specifically, the probability distributions for $\chi$ and $N$ are narrow peaked.) By applying these successive cuts, we are retaining the events with D$\chi$C formation {\it and\/} most of the pions are well-aligned along the $\pi_0$ direction. What is being cut are the events with D$\chi$C formation but most of the pions are well-aligned along the $\pi_x$ or $\pi_y$ directions, and the events where there are incoherent pions with very low $p_T$, which is the main source of noise to our signal. As a result, these successive cuts are substantially improving the signal-to-noise ratio, making it easier to observe D$\chi$C formation. On the other hand, just like any other cuts on data to suppress the noises, we are giving up on statistics. Moreover, for large $k$ we are cutting on rare events so the loss in statistics can be severe. For the cases where the signal is weak (small coherent fraction $\chi\ll 1$ or large number of domains $N\gg 1$) on each cut we are losing about two-thirds of the events. In conclusion, we have devised new cuts to enhance the signal in searches for D$\chi$C. These cuts retain only events where the $k$ pions with lowest $p_T$ are all neutral. We have shown that, after these cuts, the fraction of neutral pions within the remaining sample is substantially larger if D$\chi$Cs are formed in the heavy ion collision. \bigskip Support of this research by the U.S.~Department of Energy under grant DE-FG02-93ER-40762 is gratefully acknowledged.
\section{Introduction} The strong short-range and tensor components of the nucleon--nucleon interactions induce correlations in the nuclear wave function which are going beyond the independent--particle approximation, e.g., the Hartree--Fock method. Therefore it has always been a point of experimental and theoretical interest to find observables, which reflect these correlations in a unambiguous way. In this sense both, the overlap functions and single--nucleon spectroscopic factors, have attracted much attention in analyzing the empirical data from one--nucleon removal reactions, such as $% (e,e^{\prime }p)$, $(p,d)$, $(d,^{3}He)$ , see e.g. \cite {An93,Lap93,Bo94,Bl95} and also in other domains of many--body physics, as e.g. atomic and molecular physics \cite {Vn93,An95,Be65,Vn96,Vn97,Vn98,Sto96,Ge96}. Recently, a general procedure has been adopted \cite{Vn93} to extract the bound-state overlap functions and the associated spectroscopic factors and separation energies on the base of the ground-state one-body density matrix. The advantage of the procedure is that it avoids the complicated task for calculating the whole spectral function in nuclei. One is able instead to incorporate the knowledge of realistic one-body density matrices emerging from various correlation methods going beyond the independent--particle picture which have been proposed over the years \cite {An93,An88,Ma91,Pi92,Co92,De83,Ja55,Sa96,Fa98,Sa97,Am98,Di92,Po96,Mu95,Mut95,Ci96,Mu94}% . Initially, the procedure for extracting bound-state overlap functions has been applied \cite{Sto96} to a model one-body density matrix \cite{Sto93} accounting for the short--range nucleon correlations within the low--order approximation to the Jastrow correlation method (JCM). The calculations were based on a single harmonic--oscillator Slater determinant and Gaussian--like state--independent correlation functions. The resulting overlap functions have been used \cite{Di97} to study one--nucleon removal processes in contrast to the mean--field approaches which account for the nucleon correlations by modifying the mean--field potentials. The results obtained for the differential cross--sections of $^{16}O(p,d)$ and $^{40}Ca(p,d)$ pick--up reactions at various incident energies demonstrated that the overlap functions can be applied as realistic form factors to evaluate absolute cross--sections of such reactions. Of course, the general success of the above procedure depends strongly on the availability of realistic one-body density matrices. This work can be considered as an extension of the analysis of single--particle overlap functions based on the procedure \cite{Vn93} to more realistic one--body density matrices emerging from the correlated basis function (CBF) method \cite{Vn97,Sa96,Sa97} and the Green function method (GFM) \cite{Po96}. We have chosen the CBF theory, based on the Jastrow approach \cite{Ja55}, since it is particularly suitable for the study of the short--range correlations in nuclei. So far the calculations have been performed for infinite nuclear matter and some light nuclei as e.g. the variational Monte Carlo calculations for the $^{16}O$ nucleus \cite{Pi92}. The CBF calculations have recently been extended to medium--heavy doubly-closed shell nuclei \cite{Co92,Sa96,Sa97,Am98} using various levels of the Fermi hypernetted chain (FHNC) approximation \cite{Sa96,Fa98}. The Green function method \cite{Di92,Po96} provides detailed information on the spectral functions and nucleon momentum distributions \cite{Mu95,Mut95} predicting the largest effects of the short--range and tensor correlations at high momentum and energy \cite{Ci96,Mu94}. The main purpose of this work is twofold. Using the procedure \cite{Vn93}, we first calculate all bound-state overlap functions on the basis of one--body density matrices emerging from the CBF and Green function methods for the $^{16}O$ nucleus in order to analyze and compare their properties in coordinate and momentum spaces. Then, the resulting overlap functions are tested in the description of the experimental data from $% ^{16}O(p,d)^{15}O$ reaction for which the shape of the overlap functions is important. Such an investigation allows to examine the relationship between the one-body density matrix and the associated overlap functions within the correlation methods used and also to clarify the importance of the effects of NN correlations on the overlap functions and $(p,d)$ cross--sections. Some basic relations of the methods used to determine the effects of correlations on the one--body density matrices are given in Section II. In Section III we present numerical results for the quantities under consideration in the case of $^{16}O$. Section IV contains a summary and conclusions. \section{Correlated nuclear wave functions} The single--particle overlap functions in quantum--mechanical many--body systems are defined by the overlap integrals between eigenstates of the $A$% --particle and the $(A-1)$--particle systems: \begin{equation} \phi _{\alpha }({\bf r})=\langle \Psi _{\alpha }^{(A-1)}|a({\bf r})|\Psi ^{(A)}\rangle , \label{eq:1} \end{equation} where $a({\bf r})$ is the annihilation operator for a nucleon with spatial coordinate ${\bf r}$ (spin and isospin operators are implied). In the mean--field approximation $\Psi ^{(A)}$ and $\Psi _{\alpha }^{(A-1)}$ are single Slater determinants and the overlap functions are identical with the mean--field single--particle wave functions. Of course, this is not the case at the presence of correlations where both, $\Psi ^{(A)}$ and $\Psi _{\alpha }^{(A-1)}$, are complicated superpositions of Slater determinants. In general, the overlap functions (1) are not orthogonal. Their norm defines the spectroscopic factor \begin{equation} S_{\alpha }=\langle \phi _{\alpha }|\phi _{\alpha }\rangle . \label{eq:2} \end{equation} The normalized overlap function associated with the state $\alpha $ then reads \begin{equation} \tilde{\phi}_{\alpha }({\bf r})=S_{\alpha }^{-1/2}\phi _{\alpha }({\bf r}). \label{eq:3} \end{equation} The one--body density matrix can be expressed in terms of the overlap functions in the form: \begin{equation} \rho ({\bf r},{\bf r^{\prime }})=\sum_{\alpha }\phi _{\alpha }^{*}({\bf r}% )\phi _{\alpha }({\bf r^{\prime }})=\sum_{\alpha }S_{\alpha }\tilde{\phi}% _{\alpha }^{*}({\bf r})\tilde{\phi}_{\alpha }({\bf r^{\prime }}). \label{eq:4} \end{equation} It has been shown in \cite{Vn93} that the one-body overlap functions (\ref {eq:1}) associated with the bound states of the $(A-1)$ system can be expressed in terms of the ground state one-body density matrix of the $A$ nucleon system. In the case of a target nucleus with $J^{\pi }=0^{+}$, the lowest $(n_{0}lj)$ bound state overlap function is determined by the asymptotic behavior $(a\rightarrow \infty )$ of the corresponding partial radial contribution ${\rho _{lj}(r,r}^{\prime }{)}$ of the one-body density matrix: \begin{equation} \phi _{n_{0}lj}(r)={\frac{{\rho _{lj}(r,a)}}{{C_{n_{0}lj}~\exp (-k_{n_{0}lj}\,a})/a}}~, \label{eq:5} \end{equation} where the constants ${C_{n_{0}lj}}$ and ${k_{n_{0}lj}}$ are completely determined by ${\rho _{lj}(r,r}^{\prime }{)}$. In this way, both $\phi _{n_{0}lj}(r)$ and ${k_{n_{0}lj}}$ define the separation energy \begin{equation} \epsilon _{n_{0}lj}\equiv E_{n_{0}lj}^{(A-1)}~-~E_{0}^{(A)}=\frac{\hbar ^{2}~k_{n_{0}lj}^{2}}{2m}~ \label{eq:6} \end{equation} and the spectroscopic factor $S_{n_{0}lj}=\langle \phi _{n_{0}lj}\mid \phi _{n_{0}lj}\rangle $. The procedure also yields the next bound state overlap functions with the same multipolarity if they exist. The applicability of this procedure has been demonstrated in Refs.\cite{Vn96,Vn97,Sto96}. Thus having the procedure for estimating such important quantities as spectroscopic factors and overlap functions one has simply to apply it to some realistic one--body density matrices emerging from the CBF and Green function methods. The latter are briefly discussed in this Section. \subsection{The CBF theory} The CBF theory starts from a trial many--particle wave function \begin{equation} \Psi (x_{1},...,x_{A})={\cal S}\left[ \prod_{i<j=1}^{A}\hat{F}% (x_{i},x_{j})\right] \Phi (x_{1},...,x_{A}), \label{eq:7} \end{equation} where $A$ is the number of the nucleons with particle coordinates $% x_{1},x_{2},...,x_{A}$ which contain spatial, spin, and isospin variables, $% {\cal S}$ is a symmetrization operator, and $\Phi $ is an uncorrelated (Slater determinant) wave function normalized to unity and describing a closed-shell spherical system. The correlation factor $\hat{F}$ is generally written as \begin{equation} \hat{F}(x_{i},x_{j})=\sum_{n}h_{n}(|{\bf r}_{i}-{\bf r}_{j}|)\hat{O}_{ij}^{n} \label{eq:8} \end{equation} with basic two--nucleon operators $\hat{O}^{n}$ inducing central, spin--spin, tensor and spin--orbit correlations, either with or without isospin exchange: \begin{equation} \begin{array}{lcl} O_{ij}^{n=1,...,8} & = & 1,(\mbox{\boldmath $\tau$}_{i}.% \mbox{\boldmath $\tau$}_{j}),(\mbox{\boldmath $\sigma$}_{i}.\mbox{\boldmath $\sigma$}_{j}),(% \mbox{\boldmath $\sigma$}_{i}.\mbox{\boldmath $\sigma$}_{j})(% \mbox{\boldmath $\tau$}_{i}.\mbox{\boldmath $\tau$}_{j}),S_{ij}, \\ & & S_{ij}(\mbox{\boldmath $\tau$}_{i}.\mbox{\boldmath $\tau$}_{j}),{\bf L}.% {\bf S},{\bf L}.{\bf S}(\mbox{\boldmath $\tau$}_{i}.\mbox{\boldmath $\tau$}_{j}). \end{array} \label{eq:9} \end{equation} The corresponding one-body density matrix \begin{equation} N(x_{1},x_{1}^{\prime })=\frac{\langle \Psi |c^{\dagger }(x_{1}^{\prime })c(x_{1})|\Psi \rangle }{\langle \Psi |\Psi \rangle } \label{eq:10} \end{equation} has been calculated in \cite{Pi92} using the Monte Carlo techniques. In many cases the low--order approximation (LOA) \cite{Ga71,Da82,Fl84,Be86} is used for the ODM which consists in expanding the corresponding quantities up to the first order of the function $\hat{h}(x_{i},x_{j};x_{i}^{\prime },x_{j}^{\prime })=\hat{F} (x_{i}^{\prime },x_{j}^{\prime })\hat{F}% (x_{i},x_{j})-1$. The LOA expression for the ODM is: \begin{equation} N(x_{1},x_{1}^{\prime })=N_{0}(x_{1},x_{1}^{\prime })+N_{1}(x_{1},x_{1}^{\prime })+N_{2}(x_{1},x_{1}^{\prime }), \label{eq:11} \end{equation} with \begin{equation} N_{0}(x_{1},x_{1}^{\prime })=\rho (x_{1},x_{1}^{\prime }), \label{eq:12} \end{equation} \begin{equation} N_{1}(x_{1},x_{1}^{\prime })=\int dx_{2}\hat{h}(x_{1},x_{2};x_{1}^{\prime },x_{2})[\rho (x_{1},x_{1}^{\prime })\rho (x_{2},x_{2})-\rho (x_{1},x_{2})\rho (x_{2},x_{1}^{\prime })], \label{eq:13} \end{equation} \begin{equation} N_{2}(x_{1},x_{1}^{\prime })=\int dx_{2}dx_{3}\hat{h}% (x_{2},x_{3};x_{2},x_{3})\rho (x_{1},x_{2})[\rho (x_{2},x_{1}^{\prime })\rho (x_{3},x_{3})-\rho (x_{2},x_{3})\rho (x_{3},x_{1}^{\prime })]. \label{1eq:14} \end{equation} The zeroth order contribution (\ref{eq:12}) is the uncorrelated ODM associated with the Slater determinant $\Phi $. An important feature of the power series cluster expansion is that sum rule properties like the normalization property is fulfilled at any order of the expansion \cite{Faf98}. Thus, in our case the conservation of the number of particles, i.e., \begin{equation} \int dx_{1}N(x_{1},x_{1})=\int dx_{1}\rho (x_{1},x_{1})=A. \label{eq:15} \end{equation} is ensured. The overlap functions for $^{16}O$ have explicitly been constructed on the basis of the ODM generated by a CBF--type correlated wave function in \cite {Vn97}. The s.p. orbitals entering the Slater determinant $\Phi $ were taken from a Hartree--Fock calculation with the Skyrme--III effective force. The correlation factor $\hat{F}(x_{1},x_{2})$ obtained in \cite{Pi92} by variational calculations with Argonne NN forces has been used. The two--nucleon correlation factors were restricted to the central, spin--isospin and tensor--isospin operators. Such a description allows one to distinguish between the effects of different types of correlations on quantities such as overlap functions and spectroscopic factors of quasihole states. \subsection{FHNC formalism within the CBF theory} The CBF theory and the Fermi hypernetted chain technique have been extended in \cite{Sa96} to study medium--heavy doubly--closed shell nuclei in the $jj$ coupling scheme with different single--particle wave functions for protons and neutrons using isospin--dependent two--body correlations. These are the first microscopic calculations for nuclei beyond $^{40}Ca$ which are a necessary step towards a correct description of heavy nuclear systems based on realistic nuclear Hamiltonian. The FHNC equations can be written in terms of the one--body densities and the two--body distribution functions: \begin{equation} \rho_{1}^{\alpha }({\bf r})=\langle \Psi^{*} \sum_{k=1}^{A} \delta ({\bf r}-% {\bf r}_{k}) P_{k}^{\alpha }\Psi \rangle, \label{eq:16} \end{equation} \begin{equation} \rho_{2,q}^{\alpha \beta}=\langle \Psi^{*}\sum_{k\neq l=1}^{A} \delta ({\bf r% }-{\bf r}_{k}) P_{k}^{\alpha } \delta ({\bf r^{\prime }}-{\bf r}_{l}) P_{l}^{\beta } O_{kl}^{q}\Psi \rangle, \;\;\; q=1,...,4, \label{eq:17} \end{equation} where $P_{k}^{\alpha }$ is the projection operator on the $\alpha=p,n$ state of the $k$-nucleon. The index $q$ labels the operational component of $\rho _{2,q}^{\alpha \beta }$ with $O_{12}^{q}$ characterizing the first four channels from Eq.(\ref{eq:9}). The calculations in \cite{Sa96} have been performed using central nucleon--nucleon interactions ($v_{4}$) with spin and isospin dependence but without tensor and spin--orbit components: \begin{equation} v_{4}(1,2)=\sum_{q=1,4} v^{(q)}(r_{12})O_{12}^{q}. \label{eq:18} \end{equation} The structure of the FHNC equations depends on the adopted correlation function $f$. It has been shown in \cite{Sa96} that the isospin dependence of the correlation function within this approximation scheme is weak. This is due to the fact that within this approach the correlations mainly result from the central short--range components of the NN interaction. Therefore, in our calculations we use ODM for $^{16}O$ obtained up to the first order in the cluster expansion by adopting the Average Correlation Approximation (ACA) \cite{Sa96}. It consists in using of unique correlation, independent on the isospin of the nucleons. The ACA correlations are well reproduced by a sum of two gaussians: \begin{equation} f(r)=1- \alpha_{1} e^{-\beta_{1}r^{2}}+\alpha_{2}e^{-\beta_{2}(r-x)^{2}}, \label{eq:19} \end{equation} with the parameters: $\alpha_{1}$=0.64, $\beta_{1}$=1.54 $fm^{-2}$, $% \alpha_{2}$=0.11, $\beta_{2}$=3.51 $fm^{-2}$ and $x$=1.0 $fm$. They are taken as variational parameters fixed by minimizing the FHNC ground--state energy \cite{Sa96}. The second ingredient of these calculations is the set of s.p. wave functions which have been generated by a mean--field potential of Woods--Saxon type. The same s.p. wave functions but a state--dependent correlation function taken from nuclear matter FHNC calculations have been used in \cite{Sa97} to construct the ODM. The effects of state--dependent correlations on nucleon density and momentum distributions of various nuclei have been studied in \cite{Sa97}. It has been shown that the correlation functions used in these calculations lead to a general lowering of the density distributions in the interior region and to high--momentum components of the momentum distribution. \subsection{The Green function approach} Recent microscopic calculations of the one--body Green function for $% ^{16}O $ have demonstrated \cite{Mu95,Mut95} that the nucleon--nucleon correlations induced by the short--range and tensor components of a realistic interaction yield an enhancement of the momentum distribution at high momenta. This enhancement originates from the spectral function at large negative energies and therefore should be observed in nucleon knockout reactions with large energy transfer leaving the final nucleus at an excitation energy of about $% 100$ $MeV$. For a nucleus like $^{16}O$ with $J=0$ ground state angular momentum the one--body density matrix can easily be separated into sub--matrices of a given orbital angular momentum $l$ and total angular momentum $j$. Within the Green function approach \cite{Di92} the ODM in momentum representation can be evaluated from the imaginary part of the single--particle Green function by integrating \begin{equation} \rho _{lj}(k_{1},k_{2})=\int_{-\infty }^{\varepsilon _{F}}dE\frac{1}{\pi }% Im(g_{lj}(k_{1},k_{2};E)), \label{eq:20} \end{equation} where the energy variable $E$ corresponds to the energy difference between the ground state of the $A$ particle system and the energies of the states in the $(A-1)$--particle system (negative $E$ with large absolute value correspond to high excitation energies of the residual system) and $% \varepsilon _{F}$ is the Fermi energy. The single--particle Green function $% g_{lj}$ (or the propagator) is obtained from the solution of the Dyson equation \begin{equation} g_{lj}(k_{1},k_{2};E)=g_{lj}^{(0)}(k_{1},k_{2};E)+\int dk_{3}\int dk_{4}g_{lj}^{(0)}(k_{1},k_{3};E)\Delta \Sigma _{lj}(k_{3},k_{4};E)g_{lj}(k_{4},k_{2};E), \label{eq:21} \end{equation} where $g^{(0)}$ refers to the Hartree--Fock propagator and $\Delta \Sigma _{lj}$ represents contributions to the real and imaginary part of the irreducible self--energy, which go beyond the Hartree--Fock approximation of the nucleon self--energy used to derive $g^{(0)}$. The results for the ODM have been analyzed in \cite{Po96} in terms of the natural orbitals $\varphi_{\alpha }$ and the occupation numbers $n_{\alpha }$ in $^{16}O$ nucleus. Within the natural orbital representation they can be determined by diagonalizing the one--body density matrix of the correlated system. In this representation the radial ODM for each $lj$ subspace has the form: \begin{equation} \rho_{lj}(r,r^{\prime })=\sum_{\alpha } n_{\alpha lj}\varphi_{\alpha lj}^{*}(r) \varphi_{\alpha lj}(r^{\prime }). \label{eq:22} \end{equation} The numerical results from \cite{Po96} show that the ODM can be described quite accurately in terms of four natural orbitals $(\alpha=1,...,4)$ for each partial wave $lj$ in the sum (\ref{eq:22}). The ODM generated in this way is used in our calculations to construct the single--particle overlap functions. \section{Numerical results} The ODM obtained with the different methods mentioned in Section 2 have been applied to calculate overlap functions related to the $1s$ and $1p$ states in the $% ^{16}O$ nucleus. The ODM from \cite{Vn97,Po96} account for non-central correlation effects. Therefore one obtains in these approaches different results for $p_{3/2}$ and $p_{1/2}$ quasihole states. This allows us to calculate the corresponding differential cross--sections of $% ^{16}O(p,d)^{15}O$ reactions leading to the ground $1/2^{-}$ state and $% 3/2^{-}$ excited states of the residual $^{15}O$ nucleus. \subsection{Overlap functions and spectroscopic factors} In this subsection we present the results for the overlap functions, the spectroscopic factors and the neutron separation energies calculated using the procedure of Eqs.(\ref{eq:1})-(\ref{eq:6}) (see also \cite{Vn93}). The resulting overlap functions are compared with the HF wave function in Fig.~1. The HF wave function has been calculated in a self-consistent way using the Skyrme-III interaction. It is the uncorrelated basis function, which has also been used in \cite{Vn97}. It can be seen from Fig.~1 that the overlap functions and the HF wave function are rather similar. This is not only true for the example of $1s$ states exhibited in this figure but also for the $1p$ hole states in $^{16}O $. This justifies the use of shell--model orbitals instead of overlap functions in calculating the nucleon knock-out cross section using the Plane Wave Impulse Approximation for such kind of nuclear states. The changes in the shape from the original mean--field wave functions are rather small and might be absorbed by a suitable readjustment of the parameters of the single--particle potential used to determine the corresponding wave functions. The inclusion of short-range as well as tensor correlations leads to an enhancement of the values of the corresponding overlap functions in the interior region and a depletion in the tail region in the coordinate space. As expected, in the momentum space these effects lead to a shift of the overlap functions from the low-- to the high--momentum region in comparison with the mean--field wave functions as it is shown in Fig.~2. This fact is important because the squared overlap functions in the momentum space determine the single--particle momentum distribution representing the transition to a given single--particle state of the residual nucleus. This single--particle momentum distribution can be obtained experimentally, e.g. from $% (e,e^{\prime }p)$ reactions, by integrating the data for the spectral function over the energy interval which includes the peak of the transition. The values of the spectroscopic factors and the separation energies deduced from the calculations with different ODM are listed in Table 1. It is seen that the separation energies derived from ODM are in acceptable agreement with the corresponding single-particle energies obtained in self--consistent Hartree--Fock calculations. As it has been shown already in \cite{Vn97}, the use of single-particle wave functions which have realistic exponential asymptotics leads to separation energies of the quasihole states which are close to the original mean--field single-particle energies. The calculated spectroscopic factors, however, differ significantly from the mean--field value. The nucleon--nucleon correlations lead to a depletion of the states which are below the Fermi level of the independent particle approach. The spectroscopic factors of the $s$ and $p$ states in $% ^{16}O $ obtained within the JCM (0.94 and 0.953, respectively) \cite{Sto96} are somewhat smaller than the values obtained from the calculations which include only central channel of the interaction (about 0.98) \cite{Vn97}. Although both central correlation functions have a comparable range, the first one is more effective at small $r$, going to zero for $r=0$. Therefore the correlation effects induced by the central correlation function in this approach are stronger leading to a smaller spectroscopic factor. The same holds for the ODM generated in \cite{Sa96,Sa97}. The comparison of the spectroscopic factors also shows that the tensor correlations, which are taken into account in \cite{Vn97,Po96}, are responsible for a large part of the depletion of the occupied states. The central correlations generate a depletion of 1-2 $\%$ only, whereas the inclusion of the tensor channel leads to a depletion of 7-11 $\%$ \cite{Vn97}. The spectroscopic factors for the $% p_{3/2}$ and $p_{1/2}$ quasihole states in $^{16}O$ found in \cite{Vn97} are about 0.90--0.91. Our calculations based on the Green function theory yield similar results which indicates that about 10 $\%$ of the $1p$% --strength is removed by the short--range and tensor correlations. Here one should keep in mind that an additional depletion or reduction of the spectroscopic factors, which is not included in the approaches presented here, is due to long-range correlations\cite{skour}. In Table 1 the spectroscopic factors $S$ are given together with the natural occupation numbers $N$ \cite{An93} calculated after diagonalizing the corresponding one--body density matrices. The comparison shows that the results satisfy the general property $S_{nlj}\leq N_{nlj}^{max}$, i.e. in each $lj$ subspace the spectroscopic factor $S_{nlj}$ is smaller than the largest natural occupation number $N_{nlj}^{max}$ \cite{Vn93}. The trend of the calculated spectroscopic factors follows that of the natural occupation numbers. In the case of the Green function approach \cite{Po96,Mu95} one can compare the spectroscopic factor for the overlap function and the separation energy derived from the ODM with the occupation probability and the energy of the corresponding quasihole state listed in Table 2 of \cite{Mu95}. The occupation probabilities for the quasihole states (0.780, 0.898, 0.914 for $s_{1/2}$, $p_{1/2}$ and $p_{3/2}$, respectively) are slightly smaller than the spectroscopic factors listed in Table 1, indicating that the continuum contribution to the spectral function is non-negligible. The difference is largest for the $s_{1/2}$ state. The absolute energy of the quasihole state is slightly larger for the $s_{1/2}$ (34.3 MeV) than the corresponding separation energy (31.12 MeV) deduced from the ODM. In the case of the $p$ states the quasihole energies are smaller (14.14 MeV and 17.9 MeV) than the separation energies. \subsection{Differential cross--sections} In order to explore whether an analysis of experimental data is sensitive to the differences in the overlap functions derived from the various many-body theories, we are now going to employ a very simple model for calculating cross--sections of $^{16}O(p,d)$ reactions. The differential cross--section for such pick--up processes can be written in the form: \begin{equation} \frac{d\sigma _{pd}^{lsj}(\theta )}{d\Omega }=\frac{3}{2}\frac{S_{lsj}}{2j+1}% \frac{D_{0}^{2}}{10^{4}}\sigma _{DW}^{lsj}(\theta ), \label{eq:23} \end{equation} where $S_{lsj}$ is the spectroscopic amplitude, $j$ is the total angular momentum of the final state, $D_{0}^{2}\approx 1.5\times 10^{4}$ $MeV.fm^{3}$ is the $p-n$ interaction strength in the zero--range approximation and $\sigma_{DW}^{lsj}$ is the cross--section calculated by the DWUCK4 code \cite{La93}. For our purposes the standard Distorted Wave Born Approximation (DWBA) form factor has been replaced by the s.p. overlap function derived from the one--body density matrix calculations. In this case no extra spectroscopic factor $S_{lsj}$ in eq.(\ref{eq:23}) is needed, since our overlap functions already include the associated spectroscopic factors. The results for the differential cross--sections for the transitions to the ground $1/2^{-}$ state and to the excited $3/2^{-}$ state in $^{15}O$ nucleus at different incident proton energies $E_{p}$% =31.82, 45.34 and 65 $MeV$ are given in Figures 3--6. A comparison with the experimental data from \cite{Pre70,Ro75} is also made. The optical potentials parameter values have been taken in each case to be the same as in the corresponding standard DWBA calculations. As can be seen in Figs.3--5 the use of all overlap functions for the transition to the ground $1/2^{-}$ state leads to a qualitative agreement with the experimental data reproducing the amplitude of the first maximum and qualitatively the shape of the differential cross--section. The differences between the results obtained from the various approaches are small but visible. We emphasize that our results are without any additional normalization while the standard DWBA curves need multiplication by the fitting parameter, i.e., the spectroscopic factor. The values of the spectroscopic factors obtained from the standard DWBA procedure are given in Table 2 and can be compared with our spectroscopic factors from Table 1. It is seen that the value of the DWBA spectroscopic factors for the $1/2^{-}$ state exceeds the maximum allowed value of unity. \section{Conclusions} The results of the present work can be summarized as follows: \newline i) Single--particle overlap functions, spectroscopic factors and separation energies are calculated from the one--body density matrices, which were derived using different approximations to determine the correlated wave function for the ground state of $^{16}O$. \newline ii) The overlap functions extracted from ODM calculated within the CBF and Green function theories are peaked at smaller distance in the interior region of the nucleus compared with Hartree-Fock wave functions. \newline iii) Considering the role of the central and the tensor correlations it is found that the correlation effects on the spectroscopic factors of the hole states are dominated by the tensor channel of the interaction. \newline iv) The absolute values of the differential cross--sections of $^{16}O(p,d)$ pick--up reaction at various incident energies are calculated by using the overlap functions. The resulting angular distributions are in a qualitative agreement with the experimental cross--sections of the transitions to the ground $1/2^{-}$ and excited $3/2^{-}$ states of the residual nucleus $^{15}O$. \acknowledgments The authors are grateful to Dr. G. Co' for providing us the results for ODM from \cite{Sa96,Sa97} and to Dr. C. Giusti for the valuable discussions. This work was partly supported by the Bulgarian National Science Foundation under the Contracts Nrs.$\Phi $--527 and $\Phi $--809.
\section{Introduction} The description of exclusive nonleptonic decays of $B$ mesons represents an important and complicated theoretical problem. These decays are nonperturbative in nature and cannot be calculated reliably from the QCD Lagrangian. In contrast to exclusive semileptonic decays, where the weak current matrix elements between meson states are involved, nonleptonic decays require the evaluation of hadronic matrix elements of four local quark operators. To simplify the analysis it is usually assumed that the matrix element of the current-current weak interaction factorizes into the product of two single current matrix elements. Thus the problem reduces to the calculation of the meson form factors, which are contained in the hadronic matrix elements of weak currents as in the case of semileptonic decays, and of the meson decay constants, describing the leptonic decays. This makes the factorization hypothesis \cite{ref1} a very appealing assumption. Although it is very difficult to prove the factorization hypothesis theoretically within our present understanding of QCD nonperturbative effects, this hypothesis is expected to be valid to a rather good approximation in the case of transitions with large energy release, such as heavy $B$ decays, since the final mesons carrying large momenta escape from the region of interaction, thereby minimizing the effects of a final state interaction. Several tests have been made to prove its validity phenomenologically,\cite{ref2} and, it has been shown to work well for the description of $B$ meson decays into a $D$ or $D^*$ and a light meson. Once the factorization assumption is made, nonleptonic decays are related to the corresponding semileptonic decays. In this paper we wish to calculate the branching ratios of the exclusive nonleptonic two-meson decays of $\bar B^0 $, $B^-$ and $\bar B_s$ mesons in the framework of the covariant oscillator quark model (COQM) \cite{ref3} on the basis of factorization. One of the most important motivations of the covariant oscillator quark model (COQM) is to describe covariantly the centre of mass motion of hadrons, retaining the considerable sucesses of the non-relativistic quark model with regard to the static properties of hadrons. A key element of the COQM for achieving this is its direct treatment of the squared masses of hadrons, in contrast to the mass itself as done in conventional approaches. This makes the covariant treatment simple. The COQM has been applied to various problems \cite{ref4} with satisfactory results. Recently, Ishida et al have studied the semi-leptonic weak decays of heavy hadrons using this model \cite{ref5} and derived the same relations of weak form factors as in HQET.\cite{ref6} Furthermore, the predicted spectra for $B\rightarrow (D^*,D)l\nu$ were shown to fit experimental data quite well.\cite{spe} Keeping this success in mind, we extend application of the COQM to the nonleptonic decays of $\bar B^0 $, $B^-$ and $\bar B_s $ mesons. This paper is organized as follows. In \S 2 we present the expressions for nonleptonic decay amplitudes in the factorization approximation. In \S 3 we present a brief description of the covariant oscillator quark model. Using this model we have evaluated the form factors and obtained the decay rates for various nonleptonic decay processes. The decay modes $B \to D^*V $ are considered in \S 4. Section 5 contains our results and discussion. \section{General formalism} Neglecting the penguin contribution, the effective Hamiltonian describing the decays under consideration is given by \begin{eqnarray} &&{\cal H}_{\rm eff} = \frac{G_F}{\sqrt{2}}\; V_{cb}\;V_{q_i q_j}\; [C_1(m_b) O_1 + C_2(m_b) O_2 ] \end{eqnarray} with \begin{eqnarray} &&O_1=(\bar q_i q_j)^{\mu}\; (\bar c b)_{\mu}~~~~~~~~~~ \mbox{and}~~~~~~~~~~~ O_2= (\bar q_i b)^{\mu} (\bar c q_j)_{\mu}\;, \end{eqnarray} where $C_1$ and $C_2$ are the Wilson coefficients, and the quark current $(\bar q_i q_j )_{\mu} $ denotes the usual $(V-A)$ current. $q_i$ and $q_j $ are two types of quark flavors that are hadronized to the $P$ or $V$ mesons. The factorization approach to two-body nonleptonic decays $B \to D M $ implies that the decay amplitude can be expressed by the product of one particle matrix elements: \begin{eqnarray} \langle D M | {\cal H}_{\rm eff} | B \rangle &=& \frac{G_F}{\sqrt 2} V_{cb} V_{q_i q_j}~\biggr[ a_1 \langle D | (\bar c b)_{\mu} |B \rangle \langle M |(\bar q_i q_j )^{\mu} | 0 \rangle \nonumber\\ & & + a_2 \langle M |(\bar q_i b)_{\mu} |B \rangle \langle D |(\bar c q_j )^{\mu} | 0 \rangle \biggr] \ . \label{eq:3} \end{eqnarray} Here $a_1 = C_1 + C_2/N_c $ and $a_2=C_2 + C_1/N_c$, where $N_c$ represents the number of colors. It should be noted that in writing Eq.~(\ref{eq:3}) we have discarded the contribution of color octet currents which emerges after the Fierz rearrangement of color singlet operators. Clearly these currents violate factorization since they cannot allow transition to the vacuum states. In the factorization approximation one can distinguish\footnote{ The contributions due to quark annihilation processes, which are expected to be small,\cite{ref1} are neglected in the calculation of this paper. } three classes of $B$ meson decays : the `class I' transitions such as $\bar B^0 \to M_1^+ M_2^- $, where only the term $a_1$ contributes (both mesons produced by charged currents ); `class II' transitions, such as $ \bar B^0 \to M_1^0 M_2^0 $, where only the term $a_2$ contributes (both mesons are produced by neutral currents); and `class III' transitions, such as $B^- \to M_1^0 M_2^- $, where both terms contribute coherently. In order to evaluate the transition amplitudes we use the following matrix elements: \begin{eqnarray} &&\langle P(p) |(\bar q_i q_j)^{\mu} |0 \rangle = -if_P~ p^\mu , \nonumber\\ &&\langle V(p,\epsilon) |(\bar q_i q_j)^{\mu} |0 \rangle = M_V~f_V~ \epsilon^\mu , \nonumber\\ &&\langle a_1(p,\epsilon) |(\bar q_i q_j)^{\mu} |0 \rangle = M_{a_1}~f_{a_1}~ \epsilon^\mu\;, \label{eq:4} \end{eqnarray} where $P$, $V$ and $a_1 $ represent the pseudoscalar, vector and the axial vector mesons, respectively. To evaluate the hadronic form factors we use the COQM. These are presented in the next section. \section{ Model Framework, hadronic form factors\\ and decay width of B$\to$ PP, B$\to$ PV and B$\to$ VP} The general treatment of COQM may be called the ``boosted $LS$-coupling scheme,'' and the wavefunctions being tensors in $\tilde U(4) \times O(3,1) $-space, reduce to those in $SU(2)_{\rm spin} \times O(3)_{\rm orbit} $-space in the nonrelativistic quark model in the hadron rest frame. The spinor and space-time portion of the wave functions separately satisfy the respective covariant equations, the Bargmann-Wigner (BW) equation for the former and the covariant oscillator equation for the latter. The form of the meson wave function has been determined completely through the analysis of mass spectra. In COQM, the meson states are described by bi-local fields $\Phi_A^B(x_{1\mu},x_{2\mu}) $, where $x_{1\mu}(x_{2\mu})$ is the space time coordinate of the constituent quark (antiquark), $A=(a,\alpha)\;(B=(b,\beta))$ describing its flavor and covariant spinor. Here we write only the positive frequency part of the relevant ground state fields: \begin{equation} \Phi_A^B(x_{1\mu},x_{2\mu})=e^{iP\cdot X}\; U(P)_A^B\; f_{ab}(x_{\mu};P)\;, \end{equation} where $U$ and $f$ are the covariant spinor and internal space-time wave functions respectively, satisfying the Bargmann-Wigner and oscillator wave equations. The quantity $x_{\mu} (X_{\mu}) $ is the relative (CM) coordinate, $x_{\mu}\equiv x_{1\mu}-x_{2\mu} (X_{\mu}\equiv m_1x_{1\mu} +m_2x_{2\mu})/(m_1+m_2) $, where the $m_i$ represent the quark masses). The function $U$ is given by \begin{equation} U(P)=\frac{1}{2 \sqrt 2}\left [(-\gamma_5 P_s(v)+i\gamma_{\mu} V_{\mu}(v))(1+iv \cdot \gamma)\right ] , \end{equation} where $P_s(V_s)$ represents the pseudoscalar (vector) meson field, and $v_{\mu} \equiv P_{\mu}/M$ [$P_{\mu}(M)$ is the four momentum (mass) of the meson]. The function $U$, being represented by the direct product of quark and antiquark Dirac spinors with the meson velocity, is reduced to the non-relativistic Pauli-spin function in the meson rest frame. The function $f$ is given\footnote{ In this paper we employ the pure-confining approximation, neglecting the effect of the one-gluon-exchange potential, which is expected to be good for the heavy/light-quark meson system. } by \begin{equation} f(x_{\mu};P)=\frac{\beta}{\pi} \exp\left (-\frac{\beta}{2} \left (x_{\mu}^2+2\frac{(x\cdot P)^2}{M^2}\right ) \right )\; . \end{equation} The value of the parameter $\beta $ is determined from the mass spectra \cite{ref8} as $\beta_{\pi/\rho/a_1}=0.14 $, $\beta_{K/K^*}$ =0.142, $\beta_{D/D^*} $=0.148, $\beta_{D_s}=0.154 $, $\beta_B $ =0.151 and $\beta_{B_s} $=0.160 (in units of $ \mbox{GeV}^2$ ). The effective action for weak interactions of mesons with $W$-bosons is given by \begin{equation} S_W=\int d^4 x_1 d^4 x_2 \langle \bar \Phi_{F, P^\prime} (x_1,x_2) i\gamma_{\mu}(1+\gamma_5) \Phi_{I, P}(x_1,x_2) \rangle W_{\mu,q}(x_1)\;,\label{eq:eqn1} \end{equation} where we have denoted the interacting (spectator) quarks as 1 (2). The CKM matrix elements and the coupling constant are omitted. This equation is obtained from consideration of Lorentz covariance, assuming a quark current with the standard $V-A$ form. In Eq. (\ref{eq:eqn1}), $\Phi_{I,P}~ (\bar \Phi_{F, P^\prime}) $ denotes the initial (final) meson with definite four momentum $P_{\mu} (P_{\mu}^{\prime})$, and $q_{\mu}$ is the momentum of $W$ boson. The function $\bar \Phi $ is defined by $\bar \Phi = -\gamma_4 \Phi^\dagger \gamma_4$, and $\langle~~\rangle $ represents the trace of Dirac spinor indices. Our relevant effective current $J_{\mu}(X)_{P^\prime,P} $ is obtained by identifying the above equation with \begin{equation} S_W =\int d^4 X J_{\mu}(X)_{P^\prime,P}\;W_{\mu}(X)_q\;. \end{equation} Then $J_{\mu}(X=0)_{P^\prime,P} \equiv J_{\mu} $ is explicitly given as \cite{ptp93} \begin{eqnarray} J_{\mu} = I^{qb}(w) \sqrt{M M^\prime} &\times & [\bar P_s(v^\prime)P_s(v)(v+v^\prime )_{\mu}\nonumber\\ &+& \bar V_{\lambda}(v^\prime ) P_s(v) (\epsilon_{\mu \lambda \alpha \beta} v_{\alpha}^{\prime}v_\beta-\delta_{\lambda \mu} (w+1) - v_{\lambda}v_{\mu}^{\prime}]\;, \label{eq:10} \end{eqnarray} where $M (M^\prime) $ denotes the physical masses of the initial (final) mesons. It should be noted that in the pure confining limit, the masses of the ground state mesons are equal to the simple sums of their constituents, which are much different from the physical masses in the case of light quark pseudoscalar mesons, such as $\pi $ and $K$. Therefore we do not consider the transitions $ B \to \pi $ and $B \to K $ in our analysis as the reliabilty of the results is less for these transitions. The quantity $I^{qb}(w) $, which is the overlapping of the initial and final wave functions, represents the universal form factor. It describes the confined effects of quarks and is given by \begin{equation} I^{qb}(w)=\frac{4 \beta \beta^\prime}{\beta+\beta^\prime } \frac{1}{\sqrt{C(w)}} \exp(-G(w))\;;~~~~~~~~~~~~ C(w)=(\beta-\beta^{\prime})^2+4\beta \beta^{\prime} w^2\;, \end{equation} and \begin{eqnarray} G(w)&=& \frac{m_n^2}{2 C(w)} \biggr[ (\beta+\beta^\prime) \left \{ \left (\frac{M}{M_s} \right )^2 + \left (\frac{M^\prime} {M_s^\prime}\right )^2 -2 \frac{M M^\prime}{M_s M_s^\prime}w \right \}\nonumber\\ & & + 2 \left \{ \beta^\prime\left (\frac{M}{M_s} \right )^2 +\beta\left ( \frac{M^\prime}{M_s^\prime}\right )^2 \right \}(w^2-1) \biggr]\;, \end{eqnarray} where $M_s(M_s^\prime)$ represents the sum of the constituent quark masses of the initial (final) meson, and $m_n $ is the spectator quark mass. The form factor function $I^{cb}(w)$ for $ B \to D(D^*) $decays corresponds to the Isgur-Wise function $\xi(w) $ in HQET.\cite{ref6} At the zero recoil point $w=1$, the value of $I^{cb}(w)$ is given by \begin{equation} I^{cb}(w=1)=\frac{4 \beta \beta^\prime}{(\beta+\beta^\prime)^2}\;. \label{eq:13} \end{equation} In the heavy quark symmetry limit $\beta=\beta^\prime $, so Eq.~(\ref{eq:13}) correctly reproduces\cite{spe} the normalization condition of HQET, i.e., $\xi(w=1)=1 $. However, HQET, as it is, predicts nothing about the Isgur-Wise function except for the zero recoil point, while in COQM the form factor functions can be derived at any kinematical point of interest due to the fact that the center of mass motion of the meson there is treated covariantly, as was mentioned in \S 1. In addition, the COQM form factor $I^{qb}(w)$ is also applicable for the heavy-to-light transition processes, while HQET does not provide anything for this sector. After obtaining the effective current in the COQM, the decay widths for various $B\to D$ and $B\to D^*$ decay modes can be obtained with Eqs. (\ref{eq:3}), (\ref{eq:4}) and (\ref{eq:10}). These are as follows: \begin{eqnarray} \Gamma(B(v) &\to & D(v_1) P(v_2)) = \frac{G_F^2}{16 \pi M_{B}^2}~ ~|V_{cb} V_{q_i q_j} |^2~|{\bf p}|\nonumber\\ & \times & \biggr[a_1~f_P~\sqrt{M_{B} M_D}(I_q^{cb}(w_1)) (1+w_1)~\left (M_B-M_D\right ) \nonumber\\ & + & a_2~f_D~\sqrt{M_{B} M_P}(I_q^{qb}(w_2)) (1+w_2)~\left (M_B-M_P\right )\biggr]^2\;, \end{eqnarray} \begin{eqnarray} \Gamma(B(v) &\to & D^*(v_1) P(v_2)) = \frac{G_F^2}{16 \pi M_{B}^2}~ ~|V_{cb} V_{q_i q_j} |^2~|{\bf p}|^3\nonumber\\ &\times & \biggr[a_1~f_P~\sqrt{\frac{M_B}{ M_{D^*}}}(I_q^{cb}(w_1)) ~\left (M_B+M_{D^*}\right )\nonumber\\ &+& a_2~f_{D^*}~\sqrt{\frac{M_B}{ M_P}}(I_q^{qb}(w_2)) ~\left (M_B+M_P\right )\biggr]^2\;, \end{eqnarray} \begin{eqnarray} \Gamma(B(v) &\to & D(v_1) V(v_2)) = \frac{G_F^2}{16 \pi M_{B}^2}~ ~|V_{cb} V_{q_i q_j} |^2~|{\bf p}|^3\nonumber\\ &\times & \biggr[a_1~f_V~\sqrt{\frac{M_B}{ M_D}}(I_q^{cb}(w_1)) ~\left (M_B+M_D\right )\nonumber\\ &+& a_2~f_D~\sqrt{\frac{M_B}{ M_V}}(I_q^{qb}(w_2)) ~\left (M_B+M_V\right )\biggr]^2\;, \end{eqnarray} where we have taken $w_1 = v \cdot v_1 $ and $w_2=v \cdot v_2 $. Here $|{\bf p}| $ is the c.m. momentum of the emitted particles in the rest frame of the $B$ meson. As stated earlier, only the $a_1 $ term contributes to class I decays, and only the $a_2 $ term contributes to class II decays, while both $a_1 $ and $a_2 $ terms contribute coherently to class III decays. The COQM is also applicable to heavy to light transitions, as well as heavy to heavy transitions, such as $B\to\rho$ and $B\to K^*$. In this case the above formulas are changed by replacing $(D^*(v_1),I_q^{cb}(w_1),V_{cb})$ in $B\to D^*$ transition with $(\rho (v_1),I_q^{nb}(w_1),V_{ub})$ and $(K^*(v_1),I_q^{sb}(w_1),V_{ub})$ in $B\to\rho$ and $B\to K^*$ transitions, respectively. \section{ Decay rate, polarization and angular correlation in the decays $B \to VV $ } The helicity amplitude for the decay process $ B(p) \to D^{*}(k_1,\epsilon_1)V(k_2,\epsilon_2) $ can be expressed by three invariant amplitudes, $a,~b$ and $c$. It is given following Ref. \citen{ref9} as \begin{equation} H_{\lambda}=\epsilon_{1\mu}^{(\lambda )*} \epsilon_{2\nu}^{(\lambda )*}\left [a g^{\mu \nu} +\frac{b}{M_{D^*}M_V} p^{\mu} p^{\nu} + \frac{ic}{M_{D^*}M_V} \epsilon^{\mu \nu \alpha \beta} k_{1\alpha} p_{\beta} \right ]\;. \end{equation} The coefficients $a$, $b$ and $c$ describe the $S$-, $P$- and $D$- wave contributions to the two final vector particles. In the present framework these are given as \begin{eqnarray} a&=& \frac{G_F}{\sqrt 2} V_{cb} V_{q_i q_j} [ a_1 f_V M_V \sqrt{M_B M_{D^*}} ~I^{cb}(w_1)(1+w_1)\nonumber\\ & & ~~~~~~~~~~~~+a_2 f_{D^*} M_{D^*} \sqrt{M_B M_V} ~I^{qb}(w_2) (1+w_2) ],\nonumber\\ b&=& -\frac{G_F}{\sqrt 2} V_{cb} V_{q_i q_j}\left [ a_1 f_V M_V^2 \sqrt{\frac{ M_{D^*}}{M_B}} ~I^{cb}(w_1) +a_2 f_{D^*} M_{D^*}^2 \sqrt{\frac{M_V}{M_B}} ~I^{qb}(w_2) \right ] ,\nonumber\\ c&=&- \frac{G_F}{\sqrt 2} V_{cb} V_{q_i q_j}\left [ a_1 f_V M_V^2 \sqrt{\frac{ M_{D^*}}{M_B}} ~I^{cb}(w_1) +a_2 f_{D^*} M_{D^*}^2 \sqrt{\frac{M_V}{M_B}} ~I^{qb}(w_2) \right ] .\ \ \ \ \ \ \end{eqnarray} The helicity amplitudes are given as \begin{equation} H_{\pm 1}=a\pm \sqrt{x^2-1}~c ~~~~~~~~\mbox{and} ~~~~~~~~~H_0=-ax-b(x^2-1) \;, \end{equation} where $x$ is defined by \begin{equation} x\equiv \frac{k_1 \cdot k_2}{M_{D^*} M_V}=\frac{M_B^2 -M_{D^*}^2 -M_V^2}{2M_{D^*} M_V} \end{equation} and obeys \begin{equation} x^2 = 1+ \frac{M_B^2 |{\bf p}|^2}{M_{D^*}^2 M_V^2}\;. \end{equation} The corresponding decay rate can be obtained as \begin{equation} \Gamma(B \to D^* V)= \frac{|{\bf p}|}{8 \pi M_B^2} \biggr[2|a|^2 +|xa+(x^2-1)b|^2+2(x^2-1)|c|^2 \biggr]\;. \end{equation} The decay distribution is parametrized by the coefficients \begin{eqnarray} &&\frac{\Gamma_T}{\Gamma}=\frac{|H_{+1}|^2+|H_{-1}|^2} {|H_0|^2+|H_{+1}|^2+|H_{-1}|^2},~~~~~~~~~~~~ \frac{\Gamma_L}{\Gamma}=\frac{|H_{0}|^2} {|H_0|^2+|H_{+1}|^2+|H_{-1}|^2},\nonumber\\ &&\nonumber\\ &&\alpha_1= \frac{\mbox{Re}~(H_{+1}H_0^*+H_{-1}H_0^*)} {|H_0|^2+|H_{+1}|^2+|H_{-1}|^2},~~~~~~~~~~~~~ \beta_1=\frac{\mbox{Im}~(H_{+1}H_0^*-H_{-1}H_0^*)} {|H_0|^2+|H_{+1}|^2+|H_{-1}|^2},\nonumber\\ &&\nonumber\\ &&\alpha_2=\frac{\mbox{Re}(H_{+1}H_{-1}^*)} {|H_0|^2+|H_{+1}|^2+|H_{-1}|^2},~~~~~~~~~~~ \beta_2=\frac{\mbox{Im}(H_{+1}H_{-1}^*)} {|H_0|^2+|H_{+1}|^2+|H_{-1}|^2}. \end{eqnarray} In general, the dominant terms in the angular correlations are $\Gamma_T/\Gamma $, $\Gamma_L/\Gamma $, $\alpha_1 $ and $\alpha_2 $. The terms $\beta_1 $ and $\beta_2 $ are small since they are nonvanishing only if the helicity amplitudes $H_{+1} $, $H_{-1} $ and $H_0$ or the invariant amplitudes $a$, $b$ and $c$, respectively, have different phases. In the case of heavy to light transitions, $B\to \rho$ and $B\to K^*$, the corresponding formulas are obtained by the procedure explained as the end of the last section. \section{Results and conclusion} \begin{table} \caption{ Branching ratios of nonleptonic $\bar B^0 $ decays in the COQM. Note that the values do not include the possible contribution from the penguin diagram, which is generally expected to be of order $10^{-6}\sim 10^{-7}$. It is shown that the penguin diagram does not contribute to $\rho^+D_s^-,\rho^+D_s^{*-},\bar K^{*0}\bar D^0$ and $\bar K^{*0}\bar D^{*0}$. The contribution to $\rho^+\pi^-,\rho^+\rho^-$ and $\rho^+a_1^-$ is about $3\times 10^{-7}$.$^{1)}$ } \begin{center} \begin{tabular}{ccc} \hline \multicolumn{1}{c}{$\bar B^0 $ modes}& \multicolumn{1}{c}{This work}& \multicolumn{1}{c}{Expt. [15]}\\ \hline \multicolumn{1}{c}{Class I} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{}\\ \hline $D^+ \pi^-$& 3.00 $\times 10^{-3} $ & $(3.0\pm 0.4) \times 10^{-3} $\\ $D^+ K^-$& $ 2.28 \times 10^{-4}$ &- \\ $D^+ D^-$& $ 4.345 \times 10^{-4} $ &-\\ $D^+ D_s^-$& $ 10.9 \times 10^{-3} $ & $(8.0 \pm 3.0) \times 10^{-3} $\\ $D^+ \rho^-$& $7.406 \times 10^{-3} $ & $ (7.9 \pm 1.4) \times 10^{-3} $\\ $D^+ K^{*-}$& $ 3.84 \times 10^{-4} $ &-\\ $D^+ D^{*-}$& 3.29 $\times 10^{-4} $ & $ < 1.2 \times 10 ^{-3} $\\ $D^+ D_s^{*-}$& $0.86 \times 10^{-2} $ & $ (1.0 \pm 0.5 ) \times 10^{-2} $\\ $D^+ a_1^-$ & $6.51 \times 10^{-3} $ & $(6.0 \pm 3.3) \times 10^{-3} $\\ $D^{*+} \pi^- $ & $3.07 \times 10^{-3} $ & $(2.76 \pm 0.21 ) \times 10^{-3} $\\ $D^{*+} K^-$& $ 2.28 \times 10^{-4} $ & - \\ $D^{*+} D^-$& $3.14 \times 10^{-4} $ & $ <1.8 \times 10^{-3} $\\ $D^{*+} D_s^-$& $7.615 \times 10^{-3} $ & $(9.6 \pm 3.4) \times 10^{-3} $\\ $D^{*+} \rho^-$& $8.91 \times 10^{-3} $ & $(6.7 \pm 3.3 ) \times 10^{-3} $\\ $D^{*+} K^{*-}$& $ 4.89 \times 10^{-4} $&-\\ $D^{*+} D^{*-}$&$ 8.74 \times 10^{-4} $ & $< 2.2 \times 10^{-3} $\\ $D^{*+} D_s^{*-}$& $2.49 \times 10^{-2} $ & $(2.0 \pm 0.7) \times 10^{-2} $\\ $D^{*+}a_1^- $ & $ 0.99 \times 10^{-2} $& $(1.30 \pm 0.27 ) \times 10^{-2} $ \\ $\rho^+ D_s^{-}$ & $2.17\times 10^{-5}$ & $< 7 \times 10^{-4} $\\ $\rho^+ D_s^{*-}$ & $4.63\times 10^{-5}$ & $< 8 \times 10^{-4} $\\ $\rho^+ \pi^{-}$ & $6.53\times 10^{-6}$ & $< 8.8 \times 10^{-5} $\\ $\rho^+ \rho^{-}$ & $1.83\times 10^{-5}$ & $< 2.2 \times 10^{-3} $\\ $\rho^+ a_1^{-}$ & $1.94\times 10^{-5}$ & $< 3.4 \times 10^{-3} $\\ \hline \multicolumn{1}{c}{Class II} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} \\ \hline $D^0 \rho^0$& $ 0.649 \times 10^{-4} $ & $< 3.9 \times 10^{-4} $\\ $D^{*0}\rho^0$& $ 1.22 \times 10^{-4} $ & $ < 5.6 \times 10^{-4} $\\ $\bar K^{*0} J/\psi $ &$1.504 \times 10^{-3}$ & $(1.35 \pm 0.18) \times 10^{-3} $ \\ $\bar K^{*0} \bar D^{0}$ &$1.16 \times 10^{-6}$ & $- $ \\ $\bar K^{*0} \bar D^{*0}$ &$2.25 \times 10^{-6}$ & $- $ \\ \hline \end{tabular} \end{center} \end{table} \begin{table} \caption{ Branching ratios of nonleptonic $B^- $ decays in the COQM. The numbers for $D^0a_1^-$, $D^{*0}a_1^-$,$D^0\pi^-$,$D^{*0}\pi^-$,$D^0K^-$,$D^{*0}K^-$, $\rho^0\pi^-$ and $\rho^0a_1^-$ in class III do not contain the contributions from the color-suppressed $a_2$-term in Eq.~(2.3). Note also that the values do not include the possible contribution from the penguin diagram, which are generally expected to be of order $10^{-6}\sim 10^{-7}$. It is shown that the penguin diagram does not contribute to $\rho^0D_s^-,\rho^0D_s^{*-},\rho^-J/\psi ,K^{*-}\bar D^0$ and $K^{*-}\bar D^{*0}$. The contribution to $\rho^0\pi^-,\rho^0\rho^-$ and $\rho^0a_1^-$ is about $3\times 10^{-7}$.$^{1)}$ } \begin{center} \begin{tabular}{ccc} \hline \multicolumn{1}{c}{$B^- $ modes}& \multicolumn{1}{c}{This work}& \multicolumn{1}{c}{Expt. [15]}\\ \hline \multicolumn{1}{c}{Class I} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{}\\ \hline $D^0 D^-$& 4.53 $\times 10^{-3} $ & -\\ $D^0 D_s^-$& $ 1.16 \times 10^{-2}$ & $(1.3 \pm 0.4) \times 10^{-2} $ \\ $D^0 D^{*-}$& $ 3.48 \times 10^{-4} $ &-\\ $D^0 D_s^{*-}$& $ 9.07 \times 10^{-3} $ & $(9.0 \pm 4.0) \times 10^{-3} $\\ $D^{*0} D^-$& $3.34 \times 10^{-4} $ &-\\ $D^{*0} D_s^{-}$& 0.81 $\times 10^{-2} $ & $ ( 1.2\pm 0.5) \times 10 ^{-2} $\\ $D^{*0} D^{*-}$& $9.23 \times 10^{-4} $ & -\\ $D^{*0} D_s^{*-}$ & $2.625 \times 10^{-2} $ & $(2.7 \pm 1.0) \times 10^{-2} $\\ $\rho^0 D_s^{-}$ & 1.15 $\times 10^{-5} $ & $ < 4 \times 10 ^{-4} $\\ $\rho^0 D_s^{*-}$ & $2.45 \times 10^{-5} $ & $ < 5 \times 10^{-4} $\\ \hline \multicolumn{1}{c}{Class II} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} \\ \hline $K^{*-} J/\psi $ &$1.587 \times 10^{-3} $ & $(1.47 \pm 0.27) \times 10^{-3} $ \\ $\rho^- J/\psi $ &$5.79 \times 10^{-5} $ & $ < 7.7 \times 10^{-4} $ \\ $K^{*-} \bar D^0 $ & $1.23 \times 10^{-6} $ & $-$ \\ $K^{*-} \bar D^{*0} $ & $2.38 \times 10^{-6} $ & $-$ \\ \hline \multicolumn{1}{c}{Class III} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{} \\ \hline $D^{0} \rho^-$& $ 1.004 \times 10^{-2} $ & $ (1.34 \pm 0.18) \times 10^{-2} $ \\ $D^{*0} \rho^-$& $1.264 \times 10^{-2} $ & $(1.55 \pm 0.31) \times 10^{-2} $\\ $D^{0} a_1^-$& $6.89 \times 10^{-3} $ & $(5.0 \pm 4.0 ) \times 10^{-3} $\\ $D^{*0} a_1^{-}$& $ 1.04 \times 10^{-2} $& $(1.9 \pm 0.5 ) \times 10^{-2} $\\ $D^{0} \pi^-$& $3.18 \times 10^{-3} $ & $(5.3 \pm 0.5 ) \times 10^{-3} $\\ $D^{*0} \pi^{-}$& $ 3.25 \times 10^{-3} $& $(4.6 \pm 0.4 ) \times 10^{-3} $\\ $D^{0} K^-$& $2.41 \times 10^{-4} $ & $-$\\ $D^{0} K^{*-}$& $5.34 \times 10^{-4} $ & $-$\\ $D^{*0} K^-$& $2.42 \times 10^{-4} $ & $-$\\ $D^{*0} K^{*-}$& $7.13 \times 10^{-4} $ & $-$\\ $\rho^{0} \pi^-$& $3.45 \times 10^{-6} $ & $ < 4.3 \times 10^{-5} $\\ $\rho^{0} a_1^-$& $1.03 \times 10^{-5} $ & $ < 6.2 \times 10^{-4} $\\ $\rho^{0} \rho^-$& $1.45 \times 10^{-5} $ & $ < 1.0 \times 10^{-3} $\\ \hline \end{tabular} \end{center} \end{table} \begin{table} \caption{ Branching ratios of nonleptonic $\bar B_s^0 $ decays in the COQM model.} \begin{center} \begin{tabular}{ccc} \hline \multicolumn{1}{c}{$\bar B_s^0 $ modes}& \multicolumn{1}{c}{This work}& \multicolumn{1}{c}{Expt. [15]}\\ \hline $D_s^+ \pi^-$& 0.248 $\times 10^{-2} $ & $ < 13 \% $\\ $D_s^+ K^-$& $ 1.89 \times 10^{-4}$ &- \\ $D_s^+ D^-$& $ 3.73 \times 10^{-4} $ &-\\ $D_s^+ D_s^-$& $ 9.57 \times 10^{-3} $ &-\\ $D_s^+ \rho^-$& $6.15 \times 10^{-3} $ &-\\ $D_s^+ K^{*-}$& $ 3.196 \times 10^{-4} $ &-\\ $D_s^+ D^{*-}$& 2.86 $\times 10^{-4} $ &-\\ $D_s^+ D_s^{*-}$& $7.48 \times 10^{-3} $ &-\\ $D_s^+ a_1^-$ & $4.86 \times 10^{-3} $ & -\\ $D_s^{*+} \pi^- $ & $2.55 \times 10^{-3} $ &-\\ $D_s^{*+} K^-$& $ 1.9 \times 10^{-4} $ &-\\ $D_s^{*+} D^-$& $2.735 \times 10^{-4} $ &-\\ $D_s^{*+} D_s^-$& $6.68 \times 10^{-3} $ &-\\ $D_s^{*+} \rho^-$& $7.49 \times 10^{-3} $ &-\\ $D_s^{*+} K^{*-}$& $ 4.1 \times 10^{-4} $&-\\ $D_s^{*+} D^{*-}$&$ 6.85 \times 10^{-4} $ &-\\ $D_s^{*+} D_s^{*-}$& $2.216 \times 10^{-2} $ &-\\ $D_s^{*+}a_1^- $ & $ 8.39 \times 10^{-3} $&-\\ \hline \end{tabular} \end{center} \end{table} \begin{table} \caption{ Polarization and angular correlation parameter of $B \to D^* V $ decays.} \begin{center} \begin{tabular}{cccc} \hline \multicolumn{1}{c}{Decay modes}& \multicolumn{1}{c}{$\Gamma_L/\Gamma $}& \multicolumn{1}{c}{$\alpha_1$}& \multicolumn{1}{c}{$\alpha_2$}\\ \hline $\bar B^0 \to D^{*+} \rho^-$& 0.883 &-0.416 & 0.04 \\ $\bar B^0 \to D^{*+} D^{*-}$& 0.538 & -0.662 & 0.176\\ $\bar B^0 \to D^{*+} D_s^{*-}$& 0.515 & -0.665 & 0.187\\ $\bar B^0 \to \bar K^{*0}J/\psi $ & 0.428 & -0.605 & 0.142\\ $B^- \to D^{*0} \rho^-$& 0.855 &-0.446 & 0.044 \\ $B^- \to D^{*0} D^{*-}$& 0.538 & -0.661 & 0.176\\ $B^- \to D^{*0} D_s^{*-}$& 0.515 & -0.665 & 0.187\\ $B^- \to K^{*-}J/\psi $ & 0.428 & -0.605 & 0.141\\ \hline \end{tabular} \end{center} \end{table} In order to make the numerical estimate, we use the following values of various quantities. The quark masses (in GeV) are taken as $m_u=m_d=0.4$, $m_s=0.55$, $m_c=1.7$ and $m_b=5$.\cite{ref8} The particle masses and lifetimes are taken from Ref. \citen{ref10}. The relevant CKM parameter values used are $V_{cb}=0.0395$, $V_{cs}=1.04 $, $V_{cd}=0.224$, $V_{ud}=0.974$, $V_{us} =0.2196$ and $V_{ub}=0.08\times V_{cb}=0.00316$. The decay constants are taken as $f_{\pi}=130.7$, $f_K=159.8$; $f_{K^*}=214$;\cite{refa} $f_{\rho}=210$;\cite{refb} $f_D=220$, $f_{D^*}=230$, $f_{D_s}=240$, $f_{D_s^*}=260$\cite{refc} and $f_{a_1}=205$\cite{refd} (in MeV). The decay constant $f_{J/\psi}$ is determined from the value of $\Gamma(J/\psi \to e^+ e^- )$:\cite{ref10} \begin{equation} f_{J/\psi}=\sqrt{\frac{9}{4} \left (\frac{3}{4\pi \alpha^2} \right ) \Gamma(J/\psi \to e^+ e^-)M_{J/\psi}}=404.5 ~\mbox{MeV}\;. \end{equation} The parameters $a_1$ and $a_2$ appearing in these decays, which have recently been determined from the CLEO data,\cite{ref0} are $a_1=1.02$ and $a_2=0.23$.\cite{ref11} Using these values we obtain the branching ratios for $\bar B^0 $, $B^-$ and $\bar B_s$ mesons, which are tabulated in Tables I, II and III, respectively. The overall agreement between the predicted and experimental data is quite remarkable. Here, it may be worthwhile to note that the relevant processes are relativistic and the form factor functions $I(\omega )$ play significant roles: For example, the velocity of the final $D^+$ in the process $\bar B^0\to D^+\pi^-(\bar B^0\to D^+D_s^{*-})$ is $v_D=0.78c(0.70c)$, and the value of $I$ is $I=0.53(0.65)$. The polarization and the angular distribution parameters for $B \to D^* V$ decay modes are presented in Table IV. The polarization fraction $(\Gamma_L/\Gamma$) for $\bar B^0 \to D^{*+} \rho^- $ (0.883) agrees well with experimental value $0.93 \pm 0.05 \pm 0.05 $.\cite{ref10} For the decay mode $\bar B^0 \to \bar K^{*0}J /\psi $, $\Gamma_L/\Gamma $ (0.428) also agrees with the recent CLEO data $ 0.52 \pm 0.07 \pm 0.04 $.\cite{ref01} In this paper we have calculated the branching ratios of the exclusive nonleptonic two-meson decay of $B$ mesons using the covariant oscillator quark model based on the factorization approximation. The applied form factors are consistent with the predictions of heavy quark symmetry. The overall agreement of our predictions for two meson nonleptonic decays of $B$ mesons with the existing experimental data suggests that the factorization approximation works well and the estimation of confinement effects by the COQM is valid. \acknowledgements R. M. would like to thank CSIR, the Govt. of India, for a fellowship. A. K. G. and M. P. K. acknowledge the financial support from DST, the Govt. of India. The authors are grateful to Dr. K. Yamada for informing them of the recent values of $\beta$ obtained from the analyses of mass spectra.
\section*{I. Introduction} Haldane$^1$ in 1983 predicted that a Heisenberg antiferromagnetic (AFM) chain of integer spins has a gap in the excitation spectrum. Later, Haldane's conjecture was verified in several theoretical as well as experimental studies.$^{2-4}$ In a quasi-1D magnetic system, the integer spin chains are coupled via weak exchange interactions. If the strength of the coupling is very small, there is no possibility of ordering and individual spin chains remain in the disordered Haldane gap phase at all temperatures. In materials like $CsNiCl_{3}$, the inter-chain coupling is sufficiently strong to give rise to an AFM ordered phase below the N\'{e}el temperature $T_{N}.^{5-6}$ For temperature $T > T_{N}$, the three Haldane gap modes can be observed in experiments but as T approaches $T_{N}$, the gap vanishes at the 3D magnetic zone centre. In the ordered phase, two of the three modes become conventional gapless spin waves, while the third `longitudinal' mode develops a gap. For quasi-1D Haldane gap systems, there now exists a third possibility. Recent inelastic neutron-scattering experiments$^{7-8}$ on the mixed-spin antiferromagnets $R_2BaNiO_5$ ($R$ = $Pr$, $Nd$ or $Nd_xY_{1-x}$), provide clear evidence that the Haldane gap excitations propagating on the Ni chains survive even in the ordered phase. Unlike in systems like $CsNiCl_3$, the Haldane gap does not become zero as T approaches $T_N$. In the ordered phase, the gapped Haldane modes coexist with the conventional spin waves characteristic of the ordered phase.The Haldane gap excitations maintain their 1D character in the entire phase diagram. In this paper, we construct a model mixed spin system in 2D for which the ground state can be exactly specified. AFM long-range order exists in the ground state with gapless spin waves as excitations. The system also consists of integer spin chains, the ground states of which are the Valence Bond Solid (VBS) states. Haldane gap excitations can be created in the chains in the ordered phase of the mixed-spin system. We also discuss how the gap changes as the nature of the ordering changes. \section*{II. Ground state and excitations} Our model mixed spin system has an underlying lattice structure shown in Fig. 1. It basically consists of a square lattice at each site of which a spin-1/2 (A-spin, solid circle) is sitting. The centre of each square plaquette is occupied by a S=1 spin (B-spin, solid square). The B-spins along a chain in the horizontal direction interact via nearest-neighbour (n.n.) interactions but the individual B-spin chains do not interact with each other. The B-spins interact with the n.n. A-spins along the diagonal lines in Fig. 1. The A-spins interact with n.n. A-spins along the horizontal and vertical bonds of the square lattice. In $R_2BaNiO_5$ also, the S=1 Ni chains are decoupled and the $R^{+}_{3}$ sites with half-odd integer spins are positioned between the chains. The $R^{+}_{3}$ spins form a network via Ni-O-R and R-O-R superexchange routes. We now demand that the B-spin chains are in Valence Bond Solid (VBS) ground states. Affleck, Kenedy, Lieb and Tasaki (AKLT)$^4$ constructed a Hamiltonian for which the VBS state is the exact ground state of an integer spin chain. We follow the same principle to construct the Hamiltonian of our model. Fig. 2 shows two successive rows of A and B-spins. Consider the B-spins to be in a VBS state. Each spin-1 can be considered to be a symmetric combination of two spin-1/2's. In the VBS state, each spin-1/2 forms a singlet with a spin-1/2 at a neighbouring site. This process has to be symmetrized so that a spin-1 is restored at each site. In the ground state, a valence bond (VB) or singlet covers every link along the integer-spin chain so that the total spin of the state is zero. The Hamiltonian for which the VBS state is the exact ground state is given by \begin{equation} H_{AB}^{i} = \sum_{\Delta }^{}{P_{5/2}(\vec{S_a}+\vec{S_b}+\vec{S_c})} \end{equation} $P_{5/2}$ is the projection operator onto spin 5/2 for the triangle of spins. The sum is over successive triangles of spins with bases along the integer spin chain. The spins at sites $B_1$ and $B_2$ have magnitude 1 and the A-spin has magnitude 1/2. In the VBS state, the sum of the three spins can never be 5/2 as a VB exists between the basal spins. Thus, the projection operator $P_{5/2}$ operating on the triangle of spins in the VBS state gives zero. The VBS state is the exact ground state of $H_{AB}^i$ with eigenvalue zero. It can be easily verified that the A-spins are free when the Hamiltonian in (1) operates on the VBS state of the integer spin chain. The Hamiltonian $H_{AB}^i$, on expanding, is given by \begin{equation} H_{AB}^{i} = \frac{1}{20} [ 2 + 5 ( \vec{S_a}.\vec{S_b} + \vec{S_b}.\vec{S_c} + \vec{S_c}.\vec{S_a} ) + 2 ( \vec{S_a}.\vec{S_b} + \vec{S_b}.\vec{S_c} + \vec{S_c}.\vec{S_a} )^2 ] \end{equation} When we consider the full 2D spin system, the Hamiltonian is given by \[ H_{AB} = \sum_{i}^{}{H_{AB}^i} \] The sum is over individual B-spin chains. We now consider the interactions between the A-spins. The Hamiltonian $H_A$ is given by the usual Heisenberg AFM Hamiltonian defined on the square lattice: \begin{equation} H_A = J \sum_{<ij>}^{}{\vec{S_i}.\vec{S_j}} \end{equation} The spins have magnitude 1/2 and $<ij>$ are n.n.s. The ground state of (3), though not exactly known, has long range N\'{e}el type of order. The total Hamiltonian H is \begin{equation} H = H_A + H_{AB} \end{equation} The ground state of H has the following structure: the B-spin chains are in VBS states. The square network of A-spins is in the state which is the ground state of the S=1/2, n.n. Heisenberg AFM Hamiltonian on a square lattice. The decoupling of the S=1 spin and the S=1/2 spin ground states is possible because the A-spins are free when $H_{AB}$ operates on the VBS states. We now consider the excitations in the system. Spin wave excitations created in the A network of spins have no effect on the VBS ground states of the chains of B-spins. The spin wave excitations are gapless. Consider now the case when the network of A-spins is in the ground state and an excitation is created in a B-spin chain. Again, for convenience, we consider the two-chain strip of Fig. 2. The A-spins in the top chain are in the N\'{e}el state $\mid \uparrow \downarrow \uparrow \downarrow \uparrow \downarrow \cdots \rangle$ indicative of long range AFM order. We determine the excitation spectrum by the method of Matrix Products$^{9-11}$ The VBS ground state of the chain of B-spins can be written as \begin{equation} \mid \psi _G\rangle = Tr ( g_1\otimes g_2\otimes \cdots \otimes g_L ) \end{equation} where L is the number of sites (spins) in the chain. The $2\times 2$ matrix $g_i$ is given by \begin{equation} g_i = \left(\begin{array}{cc} -\mid 0\rangle & -\sqrt{2}\mid 1\rangle \\ \sqrt{2}\mid \bar{1}\rangle& \mid 0\rangle \end{array}\right) \end{equation} The states $\mid 1\rangle, \mid 0\rangle, \mid \bar{1}\rangle$ correspond to $S^z$ = 1, 0 and $-1$ respectively. One can easily verify that $H_{AB}\mid \psi _G\rangle = 0$. An excited state of the chain of B-spins is constructed as [3] \begin{equation} \mid \psi _{B}^{a}(k)\rangle = \frac{1}{\sqrt{L}} \sum_{j=1}^{L}{e^{-ikj}\mid \psi _{Bj}^a\rangle} \end{equation} $\mid \psi _{Bj}^a\rangle$ is obtained by replacing $g_j$ in the matrix product of the VBS state by the matrix $g_j^a$, where a = 1, 0, $ -1$ are the values of the z-component of the total spin, $S^z$, of the excited state. The matrices $g_j^a$ have the form \begin{equation} g^1 = \left(\begin{array}{cc} \sqrt{2} \mid 1\rangle& 0 \\ -\mid 0\rangle & 0 \end{array}\right), g^0 = \left(\begin{array}{cc} -\mid 0\rangle &\sqrt{2}\mid 1\rangle \\ \sqrt{2}\mid \bar{1}\rangle& -\mid 0\rangle \end{array}\right), g^{-1} = \left(\begin{array}{cc} 0 & -\mid 0\rangle \\ 0 &\sqrt{2}\mid \bar{1}\rangle \end{array}\right) \end{equation} \begin{equation} \mid \psi _{Bj}^a\rangle = Tr g_1\otimes g_2\otimes \cdots \otimes g_{j -1}\otimes g^a_j\otimes g_{j+1} \otimes \cdots \otimes g_L \end{equation} \begin{equation} \left\langle\psi _B^a(k)\mid \psi _B^a(k)\right\rangle = \frac{1}{L} \sum_{l=1}^{L}{\sum_{p=1}^{L} {e^{ik(l-p)}\left\langle\psi _{Bl}^a\mid \psi _{Bp}^a\right\rangle}} \end{equation} We now consider the case a = 1. \begin{eqnarray} \left\langle\psi _{Bl}^1\mid \psi _{Bp}^1\right\rangle = \sum_{\left\{n_\alpha ,m_\alpha \right\}}^ {}{g_{n_1n_2}^{\dag}g_{n_2n_3}^{\dag}\cdots g_{n_ln_{l+1}}^{1\dagger} g_{n_{l+1}n_{l+2}}^\dagger \cdots g_{n_Ln_1}^\dagger}\nonumber \\g_{m_1m_2}g_{m_2m_3}\cdots g^1_{m_pm_{p+1}}\cdots g_{m_Lm_1} \end{eqnarray} Define three $4\times 4$ matrices G, $G^{'}$and $G^{''}$ as \begin{eqnarray} G_{\mu _1\mu _2}& = &G_{(n_1m_1)(n_2m_2)} = g_{n_1n_2}^\dagger g_{m_1m_2}\nonumber\\ G^{'}_{\mu _1\mu _2}& =& G^{'}_{(n_lm_l)(n_{l+1}m_{l+1})} = g_{n_ln_{l+1}}^{1\dagger} g_{m_lm_{l+1}}\\ G^{''}_{\mu _1\mu _2} &=& G^{''}_{(n_pm_p)(n_{p+1}m_{p+1})} = g^\dagger _{n_pn_{p+1}} g^{1}_{m_pm_{p+1}}\nonumber \end{eqnarray} \newpage The ordering of multi-indices is given by \begin{equation} \mu = 1,2,3,4 \longleftrightarrow (11),(12),(21),(22) \end{equation} \begin{eqnarray} \left\langle\psi _{Bl}^1(k)\mid \psi _{Bp}^1(k)\right\rangle & = &\sum_{\left\{n_\alpha ,m_\alpha \right\}}^{}{G_{(n_1m_1)(n_2m_2)}G_{(n_2m_2)(n_3m_3)}}\cdots \nonumber \\ &\cdots & G^{'}_{(n_lm_l)(n_{l+1}m_{l+1})}G_{(n_{l+1}m_{l+1})(n_{l+2}m_{l+2})}\cdots\nonumber \\ &\cdots & G^{''}_{(n_pm_p)(n_{p+1}m_{p+1})}\cdots G_{(n_Lm_L)(n_1m_1)}\nonumber\\ & & \nonumber\\ &= &TrG^{l-1}G^{'}G^{p-l+1}G^{''}G^{L-p} \end{eqnarray} The matrices G, $G^{'}$ and $G^{''}$ are \begin{eqnarray} G&= &\left(\begin{array}{rrrr} 1 &0 &0 &2 \\ 0 &-1 &0 &0 \\ 0 &0 &-1 &0 \\ 2 &0 &0 &1 \end{array}\right) \nonumber \\ G^{'}& = & \left(\begin{array}{rrrr} 0 &-2 &0 &0 \\ 0 &0 &0 &0 \\ 1 &0 &0 &0 \\ 0 &-1 &0 &0 \end{array}\right) \\ G^{''}& = & \left(\begin{array}{rrrr} 0 &0 &-2 &0 \\ 1 &0 &0 & 0 \\ 0 &0 &0 &0 \\ 0 &0 &-1 &0 \end{array}\right) \nonumber \end{eqnarray} The eigenvalues and eigenvectors of G are : \begin{eqnarray} \lambda _1 = 3, &\lambda _2 = \lambda _3 = \lambda _4 = -1 & \nonumber\\ \mid e_1\rangle =\frac{1}{\sqrt{2}}\left(\begin{array}{l} 1 \\ 0 \\ 0 \\ 1 \end{array}\right) &\mid e_2\rangle = \frac{1}{\sqrt{2}}\left(\begin{array}{l} -1 \\ 0 \\ 0 \\ 1 \end{array}\right)& \\ \mid e_3\rangle =\left(\begin{array}{l} 0 \\ 1 \\ 0 \\ 0 \end{array}\right) &\mid e_4\rangle = \left(\begin{array}{l} 0 \\ 0 \\ 1 \\ 0 \end{array}\right)& \nonumber \end{eqnarray} \newpage From (14) and in the thermodynamic limit L$\to \infty $, \begin{eqnarray} \left\langle\psi _{Bl}^1(k)\mid \psi _{Bp}^1\right\rangle &=& \sum_{n=1 }^{4}{\lambda _1^{l-1}\left\langle e_1\left| G^{'}\right|n\right\rangle} \lambda _n^{p-l+1}\left\langle n \left| G^{''} \right|e_1\right\rangle \lambda _1^{L-p}\nonumber \\ & &\nonumber \\ & = &\frac{1}{2}3^L(-\frac{1}{3})^{p-l}\nonumber \end{eqnarray} Therefore, \begin{eqnarray} \left\langle\psi _B^1(k)\mid \psi _B^1(k)\right\rangle & =& \frac{1}{L}\sum_{l=1}^{L}{\sum_{p=1}^{L}{e^{ik(l-p)}\frac{1}{2}3^L (-\frac{1}{3})^{\left|p-l\right|}}}\nonumber \\ &=& 3^L\frac{2}{5+3\cos k} \end{eqnarray} The chain of A-spins is in the N\'eel state $\mid \psi _{N\acute{e}el}\rangle$. The total wave function $\mid \psi \rangle$ of the two-chain system is a product of $\mid \psi _B^1(k)\rangle$ and $\mid \psi _{N\acute{e}el} \rangle$.The norm $\left\langle\psi \mid \psi \right\rangle$ is given by the same expression as in (17). The excitation energy measured w.r.t. the ground state energy is \begin{equation} \omega _1(k) = \frac{\left\langle\psi \left|H_{AB}\right|\psi \right\rangle}{\left\langle\psi \mid \psi \right\rangle} \end{equation} The denominator has already been calculated. We now calculate the numerator. $H_{AB}$ for single chain is given by Eq. (2). The two nearest-neighbour sites of the lower chain in Fig. 2 and the intermediate site in the upper chain form the vertices of a triangle. If one of the sites of the lower chain has the `defect' matrix $g^{1}$ assosiated with it, then there are four possible states of the triangle : \begin{eqnarray} \mid \varphi _{1j}\rangle\equiv \mid\begin{array}{c} \uparrow \\ g_j^{1}\otimes g_{j+1} \end{array}\rangle, &{\large \mid}\varphi _{2j}\rangle \equiv \mid\begin{array}{c} \uparrow \\ g_j\otimes g_{j+1}^{1} \end{array} \rangle & \nonumber \\ \mid \varphi _{3j}\rangle \equiv \mid\begin{array}{c} \downarrow \\ g^{1}_j\otimes g_{j+1} \end{array}\rangle,& \mid \varphi _{4j}\rangle \equiv \mid\begin{array}{c} \downarrow \\ g_j\otimes g_{j+1}^{1} \end{array}\rangle & \end{eqnarray} One can verify that \begin{eqnarray*} H_{j,j+1}\mid \varphi _{2j}\rangle &=& 0 \\ H_{j,j+1}\mid\varphi_{4j}\rangle &=& 0 \end{eqnarray*} \begin{eqnarray} H_{j,j+1}\mid \varphi _{1j}\rangle &=& \uparrow \frac{1}{20}\left(\begin{array}{cc} -8\sqrt{2}\left(\mid 10\rangle +\mid 01\rangle\right) & -40\mid 11\rangle \\ 8\mid 00\rangle + 4\mid \bar{1}1\rangle + 4\mid 1\bar{1}\rangle& 8\sqrt{2}\left(\mid 01\rangle + \mid 10\rangle\right) \end{array}\right) \nonumber \\ &&\nonumber\\ &&+ \downarrow \frac{1}{20} \left(\begin{array}{cc} -8\mid 11\rangle & 0 \\ 4\sqrt{2}\left(\mid 01\rangle + \mid 10\rangle\right) & 8\mid 11\rangle \end{array}\right) \nonumber\\ &&\nonumber\\ H_{j,j+1}\mid \varphi _{3j}\rangle &=& \uparrow \frac{1}{20} \left(\begin{array}{cc} -8\mid 00\rangle -4\mid 1\bar{1} \rangle - 4\mid \bar{1}1\rangle & -8\sqrt{2}\left(\mid 10\rangle + \mid 01\rangle\right)\\ 4\sqrt{2}\left(\mid \bar{1}0\rangle + \mid 0\bar{1}\rangle\right)& 8\mid 00\rangle + 4\mid 1 \bar{1}\rangle + 4\mid \bar{1}1\rangle \end{array}\right) \nonumber\\ &&\nonumber\\ &&+ \downarrow \frac{1}{20}\left(\begin{array}{cc} -4\sqrt{2}\left(\mid 10\rangle + \mid 01\rangle\right) & -8\mid 11\rangle \\ 8\mid 00\rangle + 4\mid 1\bar{1}\rangle + 4 \mid \bar{1}1\rangle & 4\sqrt{2}\left(\mid 01\rangle + \mid 10\rangle\right) \end{array}\right) \end{eqnarray} The up and down spins outside the brackets represent the states of the A-spin. We represent $\mid \psi \rangle$ by the state \begin{equation} \mid \psi \rangle = \mid \begin{array}{c} \psi _{N\acute{e}el} \\ \psi _B^1(k) \end{array}\rangle \end{equation} in which the upper and lower rows contain the states of the upper and lower chains respectively. One can verify that \begin{eqnarray} & &\left\langle\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^{1}_{Bj^{'}}(k) \end{array}\left|H_{AB}\right|\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^{1}_{Bj}(k) \end{array}\right\rangle \nonumber \\ &&\nonumber\\ &=& \delta _{jj^{'}}\left\langle\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^1_{Bj}(k) \end{array}\left|\left(H_{AB}\right)_{j j+1}\right|\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^1_{Bj}(k) \end{array}\right\rangle \nonumber \\ &&\nonumber\\ &=& \delta _{jj^{'}} Tr G^{j-1}Z\left(H\right)G^{L-j-1} \end{eqnarray} where the matrix G is given in (15) and \begin{eqnarray} Z\left(H\right)_{\mu _1\mu _2} = Z\left(H\right)_{(n_1m_1)(n_2m_2)} & & \nonumber \\ = (\varphi _{1j}^\dagger )_{n_1n_2}[(H_{AB})_{j j+1}\varphi _{1j}]_{m_1m_2} & (j \,\,odd) &\nonumber \\ = (\varphi ^\dagger _{3j})_{n_1n_2}[(H_{AB})_{j j+1}\varphi _{3j}]_{m_1m_2} & (j \,\, even) & \end{eqnarray} From Eqns. (20) \begin{eqnarray} Z^{odd}(H)& =&\frac{1}{20}\left(\begin{array}{cccc} 16 &0 &0 &80 \\ 0 &-16 &0 &0 \\ 0 &0 &-16 &0 \\ 8 &0 &0 &16 \end{array}\right)\nonumber \\ Z^{even}(H)& =& \frac{1}{20}\left(\begin{array}{cccc} 8 &0 &0 &16 \\ 0 &-8 &0 &0 \\ 0 &0 & -8 &0 \\ 8 &0 &0 &8 \end{array}\right) \end{eqnarray} In the limit $L\to \infty$ , Eq. (22) reduces to \begin{eqnarray} & &\left\langle\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^1_{Bj}(k) \end{array}\left|(H_{AB})_{jj+1}\right|\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^1_{Bj}(k) \end{array}\right\rangle \nonumber \\ &&\nonumber\\ & =& \left\langle e_1 \left|G^{L-2} Z(H)\right| e_1\right\rangle\nonumber \\ &&\nonumber\\ &=& \lambda _1^{L-2} \left\langle e_1\left|Z(H)\right|e_1\right\rangle \nonumber\\ &&\nonumber\\ &=& \frac{1}{20}\times 3^{L-2}\times 60\quad for \quad j = odd\nonumber\\ &=&\frac{1}{20}\times 3^{L-2}\times 20\quad for \quad j = even \end{eqnarray} $\lambda _1$ and $\mid e_1\rangle$ are given in Eq. (16). From (7), \begin{eqnarray} & &\left\langle\psi \left|H_{AB}\right|\psi \right\rangle\nonumber\\ &=& \frac{1}{L}\sum_{j^{'}}^{}{\sum_{j}^{}{e^{ik(j^{'}-j)}\delta _{jj^{'}}\left\langle\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^1_{Bj}(k) \end{array}\left|(H_{AB})_{jj+1}\right|\begin{array}{c} \psi _{N\acute{e}el} \\ \psi ^1_{Bj}(k) \end{array}\right\rangle}}\nonumber\\ &&\nonumber\\ & =& \frac{1}{20}\times \frac{3^{L-2}}{L}(\frac{L}{2}60 + \frac{L}{2}20)\nonumber \\ &&\nonumber\\ &= &3^{L-2}\times 2 \end{eqnarray} From (18), (17) and (26), one gets \begin{equation} \omega _1(k) = \frac{1}{9}(5 + 3 \cos k ) \end{equation} The excitation energy (27) has been determined for a two-chain strip. For the 2D spin system, the chain of B-spins is connected to two chains of A-spins and the excitation energy is twice of that in (27). The excitation energy $\omega _{-1}(k)$ for $S^z = -1$ is the same as $\omega _1(k)$. The excitation energy $\omega _0(k)$ for $S^z = 0$ is $2\omega _1(k)$. Thus the gaps in the excitation spectrum for the 2D spin system are : \begin{equation} \Delta _{\pm 1} = \frac{4}{9}\quad , \quad \Delta _0 = \frac{8}{9} \end{equation} We now consider the case when excitations are created in both the A and B-chains. The excitation in the A-chain travels in the network of A-spins as a 2D spin wave. The excitation spectrum of the spin wave is gapless. The Haldane gap excitation is confined to the B-spin chain. For a ferromagnetic ordering of A-spins, one can verify that $\Delta _{+1}\neq \Delta _{-1}$, i.e., the degeneracy of the triplet Haldane modes is fully lifted. \section*{III. Conclusion} We have constructed a model mixed spin system for which the ground state can be exactly specified. The coexistence of AFM long range order and Haldane gap excitations has been explicitly shown. In the ordered phase, the degeneracy of the three Haldane gap modes is partially lifted. The transverse modes $(S^z=\pm 1)$ are still degenerate whereas the longitudinal mode $(S^z=0 )$ has an excitation gap twice that of the transverse modes. These results are in agreement with the experimental observations on the $R_2BaNiO_5$ systems.$^{7-8}$ In these systems, coexistence of gapless spin waves in an ordered phase and the Haldane gap excitations confined to Ni chains, has been observed. The three Haldane gap modes in an isolated chain of integer spins are degenerate. The lifting of the degeneracy seen in $R_2BaNiO_5$ is due to the ordering of the rare-earth spins.$^8$ It has been suggested$^{12}$ that in $R_2BaNiO_5$ the Ni-spins participate in the long range order. This is, however, not so in our model, as in the ground state there is a complete decoupling of the A and B-spins. An important feature of the $R_2BaNiO_5$ systems is the increase in the gap energy as the temperature is lowered, i.e., as the system becomes more ordered$^{12}$. Finite temperature calculations on our model are in progress and the results will be reported elsewhere. \section*{Acknowledgement} One of the Authors (EC) is supported by the Council of Scientific and Industrial Research, India under sanction No. 9/15(186)/97-EMR-I. \newpage \section*{References} $^1$F.D.M. Haldane, Phys. Lett. 93 A, 464 (1983), Phys. Rev. Lett. 50, 1153 (1983)\\ $^2$L. P. Regnault, I. Zaliznyak, J. P. Renard and C. Vettier, Phys. Rev. B 50, 9174 (1994) and references therein\\ $^3$G. F\'ath and J. S\'olyom, J. Phys.: Condens. Matter 5, 8983 (1993) and references therein\\ $^4$I. Affleck, T. Kennedy, E. H. Lieb and H. Tasaki, Phys. Rev. Lett. 59, 799 (1987); Commun. Math. Phys. 115, 477 (1988)\\ $^5$W. J. L. Buyers et al., Phys. Rev. Lett. 56, 371 (1986); R. M. Morra, W. J. L Buyers, R. L. Armstrong and K. Hirakawa, Phys. Rev. B 38, 543 (1988)\\ $^6$K. Kakurai, M. Steiner, R. Pynn and J. K. Kjems, J. Phys.: Condens. Matter 3, 715 (1991)\\ $^7$A. Zheludev, J. M. Tranquada, T. Vogt and D. J. Buttrey, Phys. Rev. B 54, 6437 (1996); Phys. Rev. B 54, 7210 (1996)\\ $^8$S. Raymond, T. Yokoo, A. Zheludev, S. E. Nagler, A. Wildes and J. Akimitsu, Cond-mat/9811040\\ $^9$A. Kl\"umper, A. Schadschneider and J. Zittartz, J.phys. A 24, L955 (1991); Z. Phys.B 87, 281 (1992); Europhys. Lett. 24, 293 (1993)\\ $^{10}$H. Niggemann and J. Zittartz, Z.Phys. B 101, 289 (1996)\\ $^{11}$K. Totsuka and M. Suzuki, J.Phys.: Condens. Matter 7, 1639 (1995)\\ $^{12}$A. Zheludev, Cond-mat/9707167 and references therein.\\ \newpage \section*{Figure Captions} Fig. 1. Model mixed spin system consisting of a square lattice of S = 1/2 spins (A-spins, solid circle). The centre of each square plaquette is occupied by a S = 1 spin (B-spin, solid square) .\\ Fig. 2. A chain of A-spins (upper row) connected to a chain of B-spins (lower row).\\ \end{document}
\section{Introduction} The study of the GeV-TeV component of gamma-ray bursts is of great importance to understand the acceleration mechanisms and the sources physical conditions. The detection of GeV gamma-rays by EGRET during some intense GRBs \cite{egret} suggests the possibility that a high energy component could be present in all events. Furthermore several models predict GeV and TeV emission, sometimes correlated with UHECRs production (see \cite{baring}, for a review). Due to the low fluxes and the small sensitive areas of satellite experiments, gamma-rays of energy larger than a few tens of GeV, must be detected by ground based experiments located at mountain altitude measuring the secondary particles generated by gamma-rays in the atmosphere. At energies E~$<$~10~TeV the number of particles reaching the ground is to small to reconstruct the shower parameters using a standard air shower array, made of several detectors spread over large areas. On the contrary, a detector consisting of a full coverage layer of counters, providing a high granularity sampling of all particle showers, can succesfully measure arrival direction and primary energy of small showers, allowing the study of the unexplored range of gamma energies between 20 GeV and 300 GeV \cite{proposal}. \section {The ARGO-YBJ detector} ARGO-YBJ is an air shower detector optimized to observe small size showers, to be constructed in the Yangbajing Laboratory (Tibet, China) at an altitude of 4300 m a.s.l. The experiment consists of a $\sim$ 71$\times$74 m$^2$ core detector realised with a single layer of RPC's ($\sim 90\%$ of active area), surrounded by an outer detector ($\sim 30\%$ of active area) for a total size of $\sim$ 100$\times$100 m$^2$. A lead converter 0.5 cm thick will cover uniformly the RPC plane in order to increase the number of charged particles by conversion of shower photons in $e^{\pm}$ and to reduce the time spread of the shower front \cite{addendum}. ARGO-YBJ can image with high efficiency and sensitivity atmospheric showers initiated by primaries of energies in the range 10~GeV$\div$~500~TeV. Its main physics goals are \cite{proposal}: Gamma-astronomy at $\sim$~100~GeV energy threshold, with a sensitivity to detect unidentified point sources of intensity as low as $10\%$ of the Crab Nebula; Gamma-Ray Burst physics, extending the satellite measurements at energies E $>$ 10 GeV; $\overline{p}/p$ ratio in the TeV energy range; Sun and heliosphere physics. The detector assembling should start in 2000 and data taking with the first $\sim 750$ $m^2$ of RPC's in 2001. \section {Sensitivity to high energy GRBs} A high energy GRB is detectable if the number of air showers from the gamma-rays is significant larger than the fluctuations of the background, due to showers from cosmic rays with arrival directions compatible with the burst position. A good angular resolution is of major importance in order to reduce the background and increase the detection sensitivity. The angular resolution and the effective area of ARGO-YBJ to detect gamma-rays as a function of the energy have been obtained by means of simulations. For gamma-rays with energy as low as E $\sim$ 10-20 GeV, the opening angle around the source direction in which 70$\%$ of the signal showers are contained is $\sim$ 5$^{\circ}$. To evaluate the sensitivity of ARGO-YBJ to detect GRBs, we considered a burst with a power law energy spectrum $dN/dE \propto E^{-\alpha}$ extending in the range 1~GeV~$\div E_{max}$, a duration $\Delta t$=1 s, and a zenith angle $\theta$=20$^{\circ}$. The burst will give a signal with a significance larger than 4 standard deviations if the energy fluence in the range 1~GeV~$\div E_{max}$ is larger than a minimum value F$_{min}$. Fig.1 shows F$_{min}$ as a function of $E_{max}$ for 3 spectral slopes. For a generic duration $\Delta t$ the minimum fluences detectable are given by $F_{min} \sqrt{\Delta t}$. \begin{figure} \vspace{0cm} \hspace{0cm}\epsfxsize=8.8cm \epsfbox{r4f1.eps} \vspace{0cm} \caption[] {The minimum energy fluence in the range 1~GeV$\div E_{max}$ of a GRB detectable by ARGO-YBJ as a function of the maximum energy of the spectrum $E_{max}$ for 3 spectral slopes.} \label{gauss} \end{figure} In the energy range considered the sensitivity is strongly dependent on the maximum energy of the spectrum $E_{max}$. ARGO-YBJ can observe GRBs with energy fluences of a few 10$^{-6}$ erg cm$^{-2}$ if the energy spectrum extends at least up to $\sim$~200~GeV with a slope $\alpha \leq$2; the minimum detectable fluence is $\sim$ 10$^{-5}$ if $E_{max}~\sim$~30~GeV. This is of particular importance, since if GRB sources are located at cosmological distances, the high energy tail of the spectrum is affected by the $\gamma \gamma \rightarrow e^+ e^-$ interaction of gamma-rays with low energy starlight photons in the intergalactic space. According to \cite{abso}, at a distance corresponding to a redshift $z$=0.1 the absorption is almost negligible, while at $z$=0.5 (1.0) the absorption becomes important for photons of energy E~$>$~100~(50)~GeV. These values give an idea of the possible maximum energy of the GRBs spectra as a function of their distance, and from Fig.1 one can infer the maximum sensitivity of ARGO-YBJ to detect cosmological GRBs. The minimum observable fluences can be compared with the fluences measured by EGRET in the 1 MeV-1 GeV energy range: $F \sim 10^{-5} \div 10^{-4}$ erg cm$^{-2}$ \cite{egret}. Since EGRET spectral slopes $\alpha$ are mostly $\sim$ 2, one could expect fluences of the same order of magnitude at energies above 1 GeV. From Fig.1 one can conclude that ARGO-YBJ could detect GRBs with the same intensity of those observed by EGRET provided that the energy spectrum extends up to few tens of GeVs; the sensitivity increases by a factor $\sim$10 for spectra extending up to $E_{max} \sim$~200~GeV.
\section*{Appendix} \def\rightarrow{\rightarrow} \def\pi^+\pi^-\gamma{\pi^+\pi^-\gamma} \def{\bf p}{{\bf p}} \defK^0{K^0} \def\bar{K^0}{\bar{K^0}} \def\alpha{\alpha} \def\bar{\alpha}{\bar{\alpha}} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\CPbar{\hbox{{\rm CP}\hskip-1.80em{/}} \newcommand{\Corrftn}[2]{C_{\vec{#1}} (\vec{#2})} \newcommand{\Clm}[3]{C_{#1 #2} (#3)} \newcommand{\Source}[2]{S_{\vec{#1}} (\vec{#2})} \newcommand{\Rftn}[2]{{\cal R}_{\vec{#1}} (\vec{#2})} \newcommand{\Rftna}[2]{{\cal R}_{\vec{#1}} (-\vec{#2})} \newcommand{\Rftnb}[2]{\bar{\cal R}_{\vec{#1}} (\vec{#2})} \newcommand{\Sourcenovec}[2]{S_{#1} (#2)} \newcommand{\Slm}[3]{S_{#1 #2} (#3)} \newcommand{\wftn}[2]{\Phi^{(-)}_{\vec{#1}} (\vec{#2})} \newcommand{\wfsquare}[2]{\left| \wftn{#1}{#2} \right|^2} \newcommand{\spectra}[1]{\frac{dN_{1}}{d\vec{#1}}} \newcommand{\etal}{{\em et al.}} \newcommand{\prc}[1]{{\em Phys.~Rev.}~C {#1}} \newcommand{\npa}[1]{{\em Nucl.~Phys.}~A {#1}} \def$\,\,$\raisebox{-.5ex}{$\stackrel{>}{\scriptstyle\sim}$$\,\,$}{$\,\,$\raisebox{-.5ex}{$\stackrel{>}{\scriptstyle\sim}$$\,\,$}} \def$\,\,$\raisebox{-.5ex}{$\stackrel{<}{\scriptstyle\sim}$$\,\,$}{$\,\,$\raisebox{-.5ex}{$\stackrel{<}{\scriptstyle\sim}$$\,\,$}} \newcommand{\ME}[3]{\left< #1 \right| #2 \left| #3 \right>} \newcommand{\bra}[1]{\left< #1 \right|} \newcommand{\ket}[1]{\left| #1 \right>} \newcommand{\ave}[1]{\left< #1 \right>} \newcommand{\dn}[2]{{d}^{#1} #2 \:} \newcommand{\dnpi}[2]{ \frac{ {d}^{#1} #2 }{ {(2\pi)}^{#1} } \:} \newcommand{\dnpiinvar}[2]{ \frac{ {d}^{#1} #2 }{2 \left|{#2}_{0}\right| {(2\pi)}^{#1} } \:} \newcommand{\dirslash}[1]{\not\!\!{#1}} \newcommand{\deltaftn}[2]{\delta^{#1}\left(#2\right)} \newcommand{\twopideltaftn}[2]{(2\pi)^{#1}\delta^{#1}\left(#2\right)} \begin{document} \title{Imaging Three-Dimensional Relative Sources from Nuclear Reactions} \author{DAVID A. BROWN} \address{Institute for Nuclear Theory, University of Washington, Box 351550\\ Seattle, WA 98195-1550, USA\\E-mail: <EMAIL>} \maketitle \abstracts{ One can access the space-time development of a heavy-ion reaction directly by imaging the source function from two particle correlation functions. In the case of like-charged pions, this imaging can be recast as a Fourier inversion problem. We will demonstrate how this inversion can be performed on full three-dimensional (i.e. in long, side and out coordinates) experimentally determined correlation functions. We will discuss the resulting three dimensional images of the relative sources. Finally, we will discuss how to perform the full three dimensional inversion for particles whose final state interactions are more complicated than those of the pions.} \section{Nuclear Interferometry} Intensity interferometry has proven to be a valuable tool for nuclear physics as it gives direct access to the space-time extent of heavy-ion reactions. The correlation function, measured in interferometry, typically is used either to extract the correlation radii or is compared to correlation functions generated from a semi-classical transport model. Recently, it was noticed that one can make better use of the correlation function -- one can use it to {\em image} the relative source function of the particles in question \cite{Brown:1997sn,Brown:1997ku}. This source function is the relative distribution of emission points of a pair of particles in their center of mass (CM) frame. The correlation radii normally extracted from the correlation functions are the widths of this distribution. While the extracted images represent an advance in the amount of information one can gather from correlation measurements, imaging was limited to angle averaged ($q_{inv}$) correlations. We now demonstrate the imaging of full three dimensional (3D) correlation functions. In this talk, we will discuss how interferometry is really an inversion problem and discuss two ways to solve the full 3D problem. Both of these methods will be applied to a simulated Coulomb corrected pion correlation function. The first method is a direct application of the Fast Fourier Transform (FFT) algorithm to the correlation data. We will discuss the various problems and limitations of this algorithm. The chief limitation of this approach is that it only works for pion pairs. The second method is a 3D implementation of the general imaging procedure used in \cite{Brown:1997sn,Brown:1997ku}. Unlike the FFT method, this method works for all particle pairs. Following the comparison of these two methods, we will outline future directions for the study of 3D correlation data. \section{The Problem} Our task is to extract as much information about the source function from the two particle correlation function as we can. The experimently measured correlation function is defined as the following ratio: \begin{equation} \Rftn{P}{q} = \Corrftn{P}{q} - 1= \frac{dN_{2}/d^2\vec{p_1}\vec{p_2}} {dN_{1}/d\vec{p_1} dN_{1}/d\vec{p_2}} - 1. \end{equation} As stated above, the source function is the relative distribution of emission points in the pair CM and in the Pratt-Koonin formalism it is \cite{Koonin:1977fh,Pratt:1990} \begin{equation} \Source{P}{r}\equiv\int \dn{}{t_1}\dn{}{t_2}\int \dn{3}{R} \;\sigma\left(\vec{R}+\vec{r}/2,t_1,\vec{P}=0\right) \sigma\left(\vec{R}-\vec{r}/2,t_2,\vec{P}=0\right). \end{equation} Here, $\sigma(\vec{r},t,\vec{p})\:d^3r\:dt$ is the probability for emitting one of the particles at time $t$ at position $\vec{r}$ with momentum $\vec{p}$. One should note that all time dependence is integrated out of the source function. For the purpose of this talk, the details of $\sigma$ are not important. We can extract the source function directly from the data because the source function and the correlation function are related through the Pratt-Koonin Equation \cite{Koonin:1977fh,Pratt:1990}: \begin{equation} \Rftn{P}{q} = \int \dn{3}{r} \left(\wfsquare{q}{r} - 1 \right) \Source{P}{r} \equiv \int \dn{3}{r} \;K(\vec{q},\vec{r})\Source{P}{r}. \label{eqn:pk} \end{equation} Here $\wftn{q}{r}$ is the pair (anti-)symmetrized wave function in the pair CM frame. One should note that this is a simple integral equation with a kernel $K(\vec{q},\vec{r})$. In the next few sections we will demonstrate the different ways to invert this integral equation. \section{Extracting the Source with the FFT} In the case of pions, the task of imaging is simple. Here, we can ignore final state interactions in the relative wavefunction, turning the problem into a Fourier transform problem. When one does this, the kernel $K(\vec{q},\vec{r})$ becomes $K(\vec{q},\vec{r}) = \cos{(2\vec{q}\cdot\vec{r})} $ and we can immediately solve for the source: \begin{equation} \Source{P}{r}=\frac{1}{\pi^3}\int\dn{3}{q} \cos{(2\vec{q}\cdot\vec{r})}\Rftn{P}{q}. \label{eqn:FFTanswer} \end{equation} This integral can then be performed with the Fast Fourier Transform algorithm. The implementation of the FFT algorithm is simple. First, we discretize eq.~(\ref{eqn:FFTanswer}), converting this equation to a finite Fourier transform. Second, we perform the resulting transform with a canned FFT routine \cite{Press:1992}. To compute the errors, we Monte Carlo sample the errors on the correlation function to generate a test correlation, invert the test correlation, then repeat. After test 100 runs, we compute the standard deviation of the ensemble of test sources. Now, while the FFT is fast, there are several problems we must deal with. The first is dealing with statistical noise. Typically this is remedied by using a filter such as the Weiner optimal filter \cite{Press:1992}. The second problem is that the data must have the number of points equal to a power of two. In other words, we need to pad data to make the next power of two, artificially increasing the resolution. The final problem is that the FFT approach only works for pion pairs. \begin{figure} \centering \includegraphics[angle=270,totalheight=2in,width=\textwidth]{corr.ps} \caption[]{Sample plot of the model correlation function. This plot is along the $q_L=q_S=0$ axis.} \label{fig:corr} \end{figure} We can now test this inversion method. To do this, we use a Gaussian test correlation with radius parameters $R_O=R_S=4$~fm and $R_L=6$~fm. We take the error bars from a real data set and then add statistical noise to the correlation to simulate an actual experimental (Coulomb corrected) correlation. Fig.~\ref{fig:corr} is a sample of this full correlation function. In fig.~\ref{fig:sources}a. we plot the results of the inversion using the FFT alone. As one can see, the tails overestimate the true source by several orders of magnitude. What is happening here is the FFT routine is Fourier transforming the statistical noise in the data. In fig.~\ref{fig:sources}b. we illustrate how the situation improves when one uses the Weiner optimal filter \cite{Press:1992}. While the tails are now much closer to the input source, they are now well below the input source. In both cases, the inverted source {\em is not consistent with the input source}. A more sophisticated filtering method could likely fix this, but given the limitations of the FFT method, it is easier to abandon it entirely. \begin{figure} \centering \includegraphics[width=\textwidth]{allsources.eps} \caption[]{Sample plots of the source functions. These plots are all along the $r_L=r_S=0$ axis. In all three panels, the dashed curve is the input source. In panel a., the source was imaged using the FFT algorithm alone. In panel b., the source was imaged using the FFT algorithm and the filter described in the appendix. In panel c., the source was imaged using the general procedure.} \label{fig:sources} \end{figure} \section{The General Approach} The authors of \cite{Brown:1997sn,Brown:1997ku} give an approach to inverting the Pratt-Koonin equation that is applicable for any particle pair and we will now outline it. We seek the source that best represents the experimental data. Making a correlation out of a test source, ${\cal R}_i^{test}-1=\sum_j K_{ij}S^{test}_j$, we would say that the test source reproduces the data well if it is the one that minimizes the $\chi^2$: \begin{equation} \chi^2=\sum_j\left[\frac{({\cal R}^{exp}(q_j)-{\cal R}^{test}(q_j))} {\Delta {\cal R}^{exp}_j}\right]^2={\rm min}. \end{equation} To find the minimum, we set $\delta \chi^2/\delta S_k = 0$ giving a matrix equation for the source: \begin{equation} S_j=[(K^{\rm T}[\Delta^2{\cal R}^{exp}]^{-1}K)^{-1}K^{\rm T} [\Delta^2{\cal R}^{exp}]^{-1}{\cal R}^{exp}]_j. \end{equation} Similarly, by performing the error analysis one finds the covariance matrix of the source: $\Delta^2S_{ij}=[K^T [\Delta^2{\cal R}^{exp}]^{-1} K]^{-1}_{ij}$. We apply this approach to the correlation described above. In fig.~\ref{fig:sources}c., we plot the source imaged this way along with the input source. One can see that this source does as good a job at reproducing the peak of the source as the FFT sources. Also, this technique does as good at reconstructing the tails as the FFT technique. However, {\em unlike} the FFT, this technique returns errors that are much more realistic. In other words, while the FFT claims it can image even below noise level, the general approach admits it can not image that well. \section*{Final Comments} We have learned several lessons by this comparison of the FFT and general approaches. While the FFT is fast it does not reliably report the uncertainty in the imaging. On the other hand, the general method does reliably estimate the uncertainty. Finally, while the FFT approach only works for pions, the general approach works for {\em any pair}. There are many questions yet to answer using the 3D sources. Do pions have a non-Gaussian tails due to resonances? Proton one dimensional sources are not Gaussian, so what do they look like in 3D? How does flow effect all 3D correlations? Given that the general method works for any pair, how about 3D unlike pair correlations? As a final note, we invite readers to download and test our one dimensional inversion code. It is available at:\\ {\centering http://www.phys.washington.edu/$\sim$dbrown/HBTprogs.html} \section*{Acknowledgments} I thank Pawe{\l} Danielewicz, Sergei Panitkin, John Cramer, Mike Lisa, and George Bertsch for discussions regarding this work. \section*{References}
\section{Introduction} The description of hadronic matter in terms of confined quark and gluon constituents carrying a color quantum number has opened the prospect of a new, deconfined, phase of matter in which colored excitations can propagate over distances much larger than typical hadronic sizes. In the framework of pure Yang-Mills theory, the transition to this new phase is thought to occur as a function of temperature. While compelling evidence for the deconfining phase transition has been collected in lattice Monte Carlo simulations \cite{mm},\cite{rothe}, it is necessary to concomitantly develop an intuitive picture for the deconfinement phenomenon in order to be able to treat scenaria as complex as heavy ion collisions; such collision experiments, planned at RHIC and LHC, are hoped to produce lumps of deconfined matter in the near future. The question of the deconfinement transition can not be separated from an underlying picture of the confinement mechanism itself. Conversely, any purported mechanism of confinement should also be able to incorporate deconfinement. The present paper concentrates on the center vortex picture of confinement in the case of SU(2) color. This mechanism, initially proposed in \cite{thoo}-\cite{corn79}, generates an area law for the Wilson loop by invoking the presence of vortices in typical configurations entering the Yang-Mills functional integral. These vortices are closed two-dimensional surfaces in four-dimensional space-time, or, equivalently, closed lines in the three dimensions making up, e.g., a time slice. They carry flux such that they contribute a factor corresponding to a nontrivial center element of the gauge group to any Wilson loop whenever they pierce its minimal area; in the case of SU(2) color to be treated below, that is a factor $-1$. If the vortices are distributed in space-time sufficiently randomly, then samples of the Wilson loop of value $+1$ (originating from loop areas pierced an even number of times by vortices) will strongly cancel against samples of the Wilson loop of value $-1$ (originating from loop areas pierced an odd number of times by vortices), generating an area law fall-off. The simplest (SU(2)) model visualization which demonstrates this is the following: Consider a universe of volume $L^4 $, and a two-dimensional slice through it of area $L^2 $, containing a Wilson loop spanning an area $A$. Generical vortices will pierce the slice at points; assume $N$ of these points to be randomly distributed on the slice. Then the probability of finding $n$ such points inside the Wilson loop area is binomial, \begin{equation} P_N (n) = \left( \begin{array}{c} N \\ n \end{array} \right) \left( \frac{A}{L^2 } \right)^{n} \left( 1 - \frac{A}{L^2 } \right)^{N-n} \end{equation} and the expectation value of the Wilson loop becomes \begin{equation} \langle W \rangle = \sum_{n=0}^{N} (-1)^{n} P_N (n) =\left( 1 - \frac{2\rho A}{N} \right)^{N} \stackrel{N\rightarrow \infty }{\longrightarrow } e^{-2\rho A} \end{equation} where the planar density of the intersection points $\rho =N/L^2 $ is kept constant as $N\rightarrow \infty $. One thus obtains an area law with string tension $\kappa = 2\rho $. In a more realistic calculation, one would e.g. take into account interactions between the vortices \cite{corr}; the proportionality constant $\kappa /\rho $ turns out to be close to $1.4$ in zero temperature lattice measurements \cite{giedt},\cite{tempv} (a survey of existing data follows further below). The emphasis of the present work, however, lies not on relatively short-range properties of the vortices such as their thickness, but on their long-range topology. This is where the argument presented above has more serious shortcomings. For one, it suggests that the expectation value of a Wilson loop might depend on the area with which one chooses to span the loop. However, due to the closed nature of the vortices, the choice of area is in fact immaterial, as it should be. In a more precise, area-independent, manner of speaking than adopted above, the value a Wilson loop takes in a given vortex configuration should be derived from the linking numbers of the vortices with the loop. Now, the above model visualization demonstrating an area law implicitly makes a strong assumption about the long-range topology of vortex configurations: For the intersection points of vortices with a given plane to be distributed sufficiently randomly on the plane to generate confinement, typical vortices or vortex networks (note that vortices are not forbidden to self-intersect) must extend over the entire universe. Consider the converse, namely that there is an upper bound to the space-time extension of single vortices or vortex networks. Then an intersection point of a vortex with a plane always comes paired with another such point a finite distance away, due to the closed character of the vortices. This pairing in particular would preclude an area law for the Wilson loop, as can be seen more clearly with the help of another simple model. Consider a universe as above, but with the additional information that intersection points of vortices with a two-dimensional slice come in pairs at most a distance $d$ apart. Then the only pairs which can contribute a factor $-1$ to a planar Wilson loop are ones whose midpoints lie in a strip of width $d$ centered on the trajectory of the loop. Denote by $p$ the probability that a pair which satisfies this condition actually does contribute a factor $-1$. This probability is an appropriate average over the distances of the midpoints of the pairs from the Wilson loop, their angular orientations, the distribution of separations between the points making up the pairs, and the local geometry of the Wilson loop up to the scale $d$. The probability $p$, however, does not depend on the macroscopic extension of the Wilson loop. A pair which is placed at random on a slice of the universe of area $L^2 $ has probability $p\cdot A/L^2 $ of contributing a factor $-1$ to a Wilson loop, where $A$ is the area of the strip of width $d$ centered on the Wilson loop trajectory. To leading order, $A=Pd$, where $P$ is the perimeter of the Wilson loop; subleading corrections are induced by the local loop geometry. Now, placing $N$ pairs on a slice of the universe of area $L^2 $ at random, the probability that $n$ of them contribute a factor $-1$ to the Wilson loop is \begin{equation} P_{N_{pair} } (n) = \left( \begin{array}{c} N_{pair} \\ n \end{array} \right) \left( \frac{pPd}{L^2 } \right)^{n} \left( 1 - \frac{pPd}{L^2 } \right)^{N_{pair} -n} \end{equation} and, consequently, the expectation value of the Wilson loop for large universes is \begin{equation} \langle W \rangle = \sum_{n=0}^{N_{pair} } (-1)^{n} P_{N_{pair} } (n) \stackrel{N_{pair} \rightarrow \infty }{\longrightarrow } e^{-\rho pPd} \end{equation} where $\rho =2N_{pair} /L^2 $ is the planar density of points. One thus observes a perimeter law, negating confinement, if the space-time extension of vortices or vortex networks is bounded. They must thus extend over the entire universe, i.e. percolate, in order to realize confinement. Conversely, therefore, a possible mechanism driving the deconfinement transition in the vortex picture is that vortices, in a sense to be made more precise below, cease to be of arbitrary length, i.e. cease to percolate, in the deconfined phase \cite{tempv}. The main result of the present work is that this is indeed the case, implying that the deconfinement transition can be characterized as a vortex percolation transition. Before entering into the details, it should be noted that a description of the deconfinement transition in terms of percolation phenomena has also been advocated in frameworks based on Yang-Mills degrees of freedom other than vortices. For one, electric flux is expected to percolate in the {\em deconfined} phase, while it does not percolate in the confined phase. Note that this is the reverse, or dual, of the magnetic vortex picture. General arguments related to electric flux percolation were recently advanced in \cite{satz}; also, specific electric flux tube models support this picture \cite{patel}. On the other hand, in the dual superconductor picture of confinement, it has been observed that the confined phase is characterized by the presence of a magnetic monopole loop percolating throughout the (lattice) universe, whereas the monopole configurations are considerably more fragmented in the deconfined phase and cease to percolate \cite{borin}. To the authors' knowledge, however, this is mainly an empirical observation and there is no clear physical argument connecting the deconfinement transition and monopole loop percolation. Indeed, there has been speculation that the two phenomena may be disconnected \cite{borin}. This should be contrasted with the vortex language, which, as discussed at length above, has the advantage of providing a clear physical picture motivating an interrelation between vortex percolation and confinement. \section{Tools and survey of existing data} Before vortex clustering properties can be investigated in detail, some technical prerequisites have to be met; foremost, one must have a manageable definition of vortices, i.e. an algorithm which allows to localize and isolate them in Yang-Mills field configurations. After the initial proposal of the center vortex confinement mechanism, a first hint of the existence of vortex configurations was provided by the Copenhagen vacuum \cite{spag} based on the observation that a constant chromomagnetic field in Yang-Mills theory is unstable with respect to the formation of flux tube domains in three-dimensional space. Later it was observed that the chromomagnetic flux associated with these domains indeed is quantized according to the center of the gauge group \cite{spagz}. However, the theory of these flux tubes quickly became too technically involved to allow e.g. the study of global properties of the flux tube networks, especially at finite temperatures. In parallel, efforts were undertaken to define and isolate vortices on a space-time lattice. One definition, proposed by Mack and coworkers \cite{mack} and developed further by Tomboulis \cite{tomold} introduces a distinction between thin and thick vortices, only the latter remaining relevant in the continuum limit. The defining property of these thick vortices is the nontrivial center element factor they contribute to a large Wilson loop when they pierce its minimal area. This definition has the advantage of being gauge invariant; on the other hand, it does not allow to easily localize vortices in the sense of associating a space-time trajectory with them. A different line of reasoning has only recently been developed in a series of papers by Del Debbio et al \cite{giedt},\cite{deb96}-\cite{deb97aug}. One chooses a gauge which as much as possible concentrates the information contained in the field configurations on particular collective degrees of freedom, in the present case, the vortices. If this concentration of information is successful (more about this question further below), one obtains a good approximation of the dynamics by neglecting the residual deviations away from the chosen collective degrees of freedom, i.e. by projecting onto them. This type of approach was pioneered by G.~'t~Hooft, who introduced the class of Abelian gauges and the subsequent Abelian projection in order to study Abelian monopole degrees of freedom \cite{tho81}. In complete analogy, one can introduce maximal center gauges \cite{giedt},\cite{deb96}-\cite{deb97aug}, in which one uses the gauge freedom to choose link variables on a space-time lattice as close as possible to center elements of the gauge group. Subsequently, one can perform center projection, i.e. replace the gauge-fixed link variables with the center elements nearest to them on the group. Given such a lattice of center elements, i.e. in the case of SU(2) color, a lattice with links taking the values $\pm 1$, center vortices are defined as follows: Consider all plaquettes in the lattice. If the links bordering the plaquette multiply to $-1$, then a vortex pierces that plaquette. These are precisely the vortices needed for the center vortex mechanism of confinement. To see this, one merely needs to apply Stokes' theorem: Consider a Wilson loop $W$, made up of links $l=\pm 1$, and an area $A$ it circumscribes, made up of plaquettes $p=\pm 1$ (the value of a plaquette is given by the product of the bordering links). Then \begin{equation} W = \prod_{l \in W} l = \prod_{p \in A} p \end{equation} (the same letter was used here to denote both space-time objects and the associated group elements). In other words, the Wilson loop receives a factor $-1$ from every vortex piercing the area. Furthermore, the product of all plaquettes making up a three-dimensional elementary cube in the lattice is $1$, since this product contains every link making up the cube twice. This fact, which in physical terms is a manifestation of the Bianchi identity, implies that every such cube has an even number of vortices piercing its surfaces; consequently, any projection of the lattice down to three dimensions contains only closed vortex lines. Since any cut through a two-dimensional vortex surface in four dimensions is thus a closed line, the original surface is also closed. Note that if one defines the dual lattice as a lattice with the same spacing $a$ as the original one, shifted with respect to the latter by the vector $(a/2,a/2,a/2,a/2)$, then vortices are made up of plaquettes on the dual lattice. In the work presented here, the specific maximal center gauge called ``direct maximal center gauge'', see e.g. \cite{giedt}, was used. This gauge is reached by maximizing the quantity \begin{equation} \sum_{l} \left| \mbox{tr} \, U_l \right|^{2} \ , \end{equation} where $l$ labels all the links $U_l $ on the lattice. Center projection then means replacing \begin{equation} U_l \rightarrow \mbox{sign tr } U_l \ . \end{equation} In practice, the question whether the gauge fixing and projection procedure indeed successfully concentrates the relevant physical information on the collective degrees of freedom being projected on is difficult to settle a priori; most often, this is tested a posteriori by empirical means. Success furthermore depends on the specific physics, i.e. the observable, under consideration. One carries out two Monte Carlo experiments, using the full Yang-Mills action as a weight in both cases, and samples the observable in question, such as e.g. the Wilson loop, using either the full lattice configurations or the center projected ones. If the results agree, one refers to this state of affairs as ``center dominance'' for that particular observable. Center dominance for the Wilson loop is interpreted as evidence that the center gauge concentrates the physical information relevant for confinement on the vortex degrees of freedom, and that consequently center projection, i.e. projection onto the associated vortex configuration, constitutes a good approximation. Center dominance has been verified for the long-range part of the static quark potential at zero temperature (see \cite{deb96}-\cite{deb97aug} for the SU(2) theory and \cite{giedt} for the SU(3) theory). This recent verification of center dominance has sparked renewed interest in the vortex picture of confinement. In establishing the relevance of vortex degrees of freedom for confinement, it provides the necessary basis for any further investigation of vortex properties. An observation analogous to center dominance has been made in the framework of the gauge-invariant vortex definition advanced by Tomboulis \cite{tom97}. There, one samples both the quantities $W$ and sign$(W)$, $W$ denoting the Wilson loop; sign$(W)$ is interpreted as containing only the center vortex contributions to $W$, whereas all other fluctuations of the gauge fields are neglected. One finds that the expectation value of sign$(W)$ alone already provides the full string tension, i.e. one finds a gauge-invariant type of center dominance (see \cite{tom97} for the SU(2) theory and \cite{tom98} for the SU(3) theory). Subsequently, it has been noted that this type of center dominance without gauge fixing can in fact be understood in quite simple terms \cite{ogilv}, and that furthermore the density of center vortices arising on center-projected lattices without gauge fixing does not exhibit the renormalization group scaling corresponding to a finite physical density \cite{faunf}. In parallel, other vortex properties were investigated. There is evidence in the SU(2) theory that the vortices defined by center gauging and center projection indeed localize thick vortices as defined by their center element contributions to linked Wilson loops \cite{deb97},\cite{deb97aug}. In both the gauge-fixed and unfixed frameworks, absence of vortices was shown to imply absence of confinement \cite{deb97},\cite{deb97aug},\cite{tomabs}. In zero temperature lattice calculations using the maximal center gauge, the planar density of intersection points of vortices with a given surface was shown to be a renormalization group invariant, physical quantity in the SU(2) theory, cf. \cite{kurt} (note erratum in \cite{tempv}) and also \cite{giedt}. This planar density equals approximately $3.6/\mbox{fm}^{2} $ if one fixes the scale by positing a string tension of $(440 \mbox{MeV})^{2} $. Also the radial distribution function of these intersection points on a plane is renormalization group invariant \cite{corr}. Furthermore, if one takes into account the thickness of center vortices, they are able to account for the ``Casimir scaling'' behavior of higher representation Wilson loops, a feature which hitherto was considered incompatible with the vortex confinement mechanism \cite{fa97},\cite{fa98}. Also, the monopoles generated by the maximal Abelian gauge have been found to lie on the center vortices identified in a subsequent (indirect) maximal center gauge, forming monopole-antimonopole chains \cite{deb97aug}. Recently, a modified $SU(2)$ lattice ensemble was investigated in which all center vortices had been removed, with the result that chiral symmetry is restored and all configurations turn out to belong to the topologically trivial sector \cite{phil}. The purpose of the present analysis is to confront the center vortex picture of confinement with the finite temperature transition to a deconfined phase observed in Yang-Mills lattice experiments. Some previous work on vortex properties at finite temperatures has already been carried out, generalizing the zero-temperature results surveyed above. For one, the authors reported some preliminary work in \cite{tempv}. There, center dominance for the string tension between static quarks was verified at finite temperatures, and the transition to the deconfined phase with a vanishing string tension observed at the correct temperature in the center-projected theory. A depletion in the density of vortex intersection points with a plane extending in the (Euclidean) time and one space direction occurs as one crosses into the deconfined phase. The vortices are to a certain extent polarized in the time direction. However, the polarization is not complete; an area spanned by a Polyakov loop correlator is still pierced by a finite density of vortices. Thus, more detailed correlations between these vortex intersection points must induce the deconfinement transition; this led the authors to first conjecture in \cite{tempv} that the deconfinement transition in the center vortex picture may be connected to global properties of vortex networks such as their connectivity. Very recently, a related investigation into the global topology of the two-dimensional vortex surfaces in four space-time dimensions was reported in \cite{bertl}, including the case of finite temperatures. This investigation focused on properties such as orientability and genus of the surfaces, in particular, changes in these characteristics as one crosses into the deconfined phase. In the present work, the global properties of vortex surfaces are considered from a slightly different vantage point, namely specifically with a view to testing the heuristic arguments given in the introduction, connecting confinement with percolation properties. For this purpose, it will be necessary to consider in more detail different slices of vortex surfaces; details follow below. \section{Spatial string tension} \label{centdo} Before doing so, a certain gap in the existing literature on center vortices at finite temperatures should be addressed. As already mentioned above, the basis for the center vortex picture of confinement is the empirical observation of {\em center dominance} for the Wilson loop. Without first establishing center dominance for an observable under investigation, a more detailed discussion of the manner in which vortex dynamics influence the observable runs the risk of being largely academic. Center dominance for the finite-temperature long-range heavy quark potential, via the corresponding Polyakov loop correlator, was verified in \cite{tempv}, as mentioned above; however, in what follows, also the behavior of the so-called spatial string tension, extracted from large spatial Wilson loops, will be under scrutiny. To provide the necessary basis for this, center dominance for large spatial Wilson loops should first be checked. For this purpose, the authors have carried out lattice measurements of spatial Creutz ratios, using center-projected configurations to evaluate the Wilson loops, for three temperatures. Before presenting the results, a comment on the physical scales is in order. Throughout this paper, the zero-temperature string tension is taken to be $\kappa = (440 \mbox{MeV})^2 $, the lattice spacing $a(\beta ) $ at inverse coupling $\beta = 2.3$ is determined by $\kappa a^2 =0.12$, and one-loop scaling is used for the $\beta $-dependence of $a$. The deconfinement temperature is identified as $T_C =300$ MeV, cf. \cite{tempv}. It should be noted that these scales are fraught with considerable uncertainty, of the order of $10\% $, due to finite size effects. This was discussed in more detail in \cite{tempv}. The values obtained for the center-projected spatial Creutz ratios are summarized in Fig.~\ref{figa}, where they are compared with the high-precision data for the full spatial string tension of Bali et al \cite{karsch}. Since the temperatures used here and in \cite{karsch} do not coincide, an interpolation of the data points given in \cite{karsch} had to be carried out to arrive at the values depicted in Fig.~\ref{figa}. \begin{figure} \vspace{-4cm} \centerline{ \epsfysize=14cm \epsffile{spst.ps} } \caption{Center-projected spatial Creutz ratios in units of the zero-temperature string tension, $\kappa (T\! \! =\! \! 0) a^2 $; $l\times l$ Creutz ratios are displayed at $r=\sqrt{l(l-1)} a(\beta )$ on the horizontal axis. Measurements were taken on a $12^3 \times N_t $ lattice. Shown are the temperatures $T=1.1T_C $ (open symbols; triangles pointing left correspond to $\beta =2.32, N_t =4$, whereas triangles pointing right correspond to $\beta =2.4, N_t =5$), $T=1.4T_C $ (filled symbols; diamonds correspond to $\beta =2.4, N_t =4$, whereas circles correspond to $\beta =2.3, N_t =3$), and $T=1.7T_C $ (crosses correspond to $\beta =2.48, N_t =4$, whereas 'x's correspond to $\beta =2.37, N_t =3$). For comparison, the spatial string tension extracted from full Wilson loops, as interpolated from data reported by Bali et al \cite{karsch}, is displayed: The triangle pointing downwards corresponds to $T=1.4T_C $, whereas the triangle pointing upwards corresponds to $T=1.7T_C $. The full spatial string tension at $T=1.1T_C $ is virtually indistinguishable from the zero-temperature value.} \label{figa} \end{figure} Measurements were taken on a $12^3 \times N_t $ lattice, and for each temperature, two values of the inverse coupling $\beta $ were used. Note that there are two potential sources of scaling violations in Fig.~\ref{figa}. On the one hand, center projection may destroy the renormalization group scaling of the spatial string tension known to occur when using the full configurations \cite{karsch}. This type of scaling violation would be a consequence, and thus a genuine indicator, of vortex physics. On the other hand, the manner in which the data is presented in Fig.~\ref{figa} also engenders additional scaling violations to the extent in which Creutz ratios, which represent difference quotients with increment $a(\beta )$, still deviate from the derivatives they converge to as $a\rightarrow 0$. The authors have elected to accept this slight disadvantage, since the presentation of the data in Fig.~\ref{figa} is on the other hand well adapted to aid in the discussion below. Now, comparing the data obtained for different $\beta $ at one temperature in Fig.~\ref{figa}, scaling violations are evidently not significant as compared with the error bars. Namely, values of Creutz ratios for two different choices of $\beta $ are well described by a universal curve, better in fact than the error bars would suggest. However, in view of the size of the error bars, which is due to the moderate statistics available to the authors, the data do not give very stringent evidence of correct renormalization group scaling; they are perhaps best described as being compatible with such scaling. Furthermore, the data seem to point towards a certain change in the dynamics generating the spatial string tension as the temperature is raised to values significantly above the deconfinement transition. At $T=1.1T_C $, the Creutz ratios are practically constant as a function of the Wilson loop size. This behavior of center-projected Wilson loops has been reported before in zero-temperature studies \cite{deb97} and has been dubbed ``precocious scaling''. Center projection truncates the short-range Coulomb behavior of full Wilson loops and one can read off the asymptotic string tension already from $2\times 2$ Creutz ratios. By contrast, this behavior does not seem quite as pronounced at temperatures significantly above the deconfinement transition. Creutz ratios rise as a function of loop size; it should however be mentioned that this rise is much weaker than the usual Coulomb fall-off one obtains when using the full Yang-Mills configurations to evaluate the Creutz ratios. Due to this variation with loop size, the asymptotic value of the full spatial string tension extracted from the data in \cite{karsch} is, in the case of $T=1.4T_C $, only reached by the Creutz ratio corresponding to the largest Wilson loops investigated; at $T=1.7T_C $, the asymptotic value is not quite reached even by the ratios derived from the most extended loops sampled, although it is within the error bars. While the error bars afflicting the Creutz ratios extracted from larger loops are sizeable, the rise as a function of loop size does seem to be significant, especially as compared to the precocious scaling displayed at $T=1.1T_C$. Also the difference between the values taken at $T=1.4T_C $ and $T=1.7T_C $ is compatible with the difference found for the full Wilson loops \cite{karsch}. In view of their limited accuracy, the data depicted in Fig.~\ref{figa} are perhaps best done justice by the statement that they do not allow to negate the hypothesis of center dominance for the spatial string tension in the deconfined phase. Certainly, no drastic deviation from center dominance is apparent. However, more accurate studies of this question are clearly called for. \section{Vortex percolation} \subsection{Clustering of vortices} As already mentioned above, there exists even in the deconfined phase a substantial density of vortex intersection points on the area spanned by two Polyakov loops \cite{tempv}. Thus, deconfinement must be due more specifically to a correlation between these intersection points, such that the distribution of points ceases to be sufficiently random to generate an area law. As motivated in the introduction, a correlation conducive to deconfinement would occur if vortices only formed clusters smaller than some maximal size, i.e. if they ceased to percolate. This would make the points appear in pairs separated by less than the aforementioned maximal size, leading to a perimeter law for the Polyakov loop correlator. In order to test whether this type of mechanism is at work in connection with the Yang-Mills deconfinement transition, it is necessary to measure the extension of vortex clusters. Vortices constitute closed two-dimensional surfaces in four space-time dimensions, or, equivalently, one-dimensional loops if one projects down to three dimensions by taking a fixed time slice or a fixed space slice of the (lattice) universe. Note that the term space slice here is meant to denote the three-dimensional space-time one obtains by holding just one of the three space coordinates fixed. Which particular coordinate is fixed is immaterial in view of spatial rotational invariance. In the following, specifically the extension of vortex line clusters in either time or space slices will be investigated. In this way, the relevant information is exhibited more clearly than by considering the full two-dimensional vortex surfaces in four-dimensional space-time. Given a center-projected lattice configuration, the corresponding vortices can be constructed on the dual lattice in the fashion already indicated in the introduction. As a definite example, consider a fixed time slice. Then the vortices are described by lines made up of links on the dual lattice. Consider in particular a plaquette on the original lattice, lying e.g. in the $z=z_0 $ plane and extending from $x_0 $ to $x_0 +a$ and from $y_0 $ to $y_0 +a$, where $a$ denotes the lattice spacing. By definition, if the links making up this plaquette multiply to the center element $-1$, then a vortex pierces that plaquette. This means that a certain link on the dual lattice is part of a vortex; namely, the link connecting the dual lattice points $(x_0 +a/2 , y_0 +a/2 , z_0 -a/2 )$ and $(x_0 +a/2 , y_0 +a/2 , z_0 +a/2 )$. Having constructed the vortex configuration on the dual lattice, one can proceed to define the vortex clusters. One begins by scanning the dual lattice for a link which is part of a vortex. Starting from that link, one tests which adjacent links, i.e. links which share a dual lattice site with the first link, are also part of the vortex. This is repeated with all new members of the cluster until all links making up the cluster are found. In this way, it is possible to separate the different vortex clusters. \subsection{Extension of vortex clusters} \label{secsp} Given the vortex clusters, their extensions can be measured. Consider all pairs of links on a cluster and evaluate the space-time distance between each pair. The maximal such distance defines the extension of that cluster. In Figs.~\ref{fig2}-\ref{fig4}, histograms are displayed in which, for every cluster, the total number of links making up that cluster was added to the bin corresponding to the extension of the cluster. \begin{figure}[h] \centerline{ \hspace{0.5cm} \epsfysize=8cm \epsffile{scl0.7.ps} \hspace{0.5cm} \epsfysize=8cm \epsffile{scl0.8.ps} } \caption{Vortex material distributions in space slices of $12^3 \times N_t $ lattice universes obtained as described in the text. Left: $12^3 \times 8$ lattice at $\beta =2.4$, which is identified with $T=0.7T_C $. Right: $12^3 \times 7$ lattice at $\beta =2.4$, which is identified with $T=0.8T_C $. The bins represent the percentage of vortex material organized into clusters of the corresponding extension. Extension is measured on the horizontal axis in units of the maximal extension possible on a space slice of the given lattice, namely $\sqrt{2\cdot (12/2)^2 + (N_t /2)^2 } $ lattice spacings.} \label{fig2} \end{figure} \begin{figure}[h] \centerline{ \hspace{0.5cm} \epsfysize=8cm \epsffile{scl0.9.ps} \hspace{0.5cm} \epsfysize=8cm \epsffile{scl1.1.ps} } \caption{Vortex material distributions as in Fig.~\ref{fig2}, at different temperatures. Left: $12^3 \times 6$ lattice at $\beta =2.4$, which is identified with $T=0.9T_C $. Right: $12^3 \times 5$ lattice at $\beta =2.4$, which is identified with $T=1.1T_C $.} \label{fig2a} \end{figure} \begin{figure}[h] \centerline{ \hspace{0.5cm} \epsfysize=8cm \epsffile{scl1.4.ps} \hspace{0.5cm} \epsfysize=8cm \epsffile{scl1.8.ps} } \caption{Vortex material distributions as in Fig.~\ref{fig2}, at different temperatures. Left: $12^3 \times 4$ lattice at $\beta =2.4$, which is identified with $T=1.4T_C $. Right: $12^3 \times 3$ lattice at $\beta =2.4$, which is identified with $T=1.85T_C $.} \label{fig3} \end{figure} The histograms were finally normalized such that the integral of the distributions gives unity. Constructed in this way, the histograms give a very transparent characterization of typical vortex configurations. The content of each bin represents the percentage of the total vortex length in the configurations, i.e. the available vortex material, which is organized into clusters of the corresponding extension. Accordingly, these distributions will be referred to as {\em vortex material distributions} in the following. In a percolating phase, the vortex material distribution is peaked at the largest extension possible on the lattice universe under consideration. Note that, due to the periodic boundary conditions, this maximal extension e.g. on a $N_s \times N_s \times N_t $ space slice of the four-dimensional space-time lattice is $\sqrt{(N_s /2)^2 + (N_s /2)^2 + (N_t /2)^2 } $ lattice spacings. In a non-percolating phase, the vortex material distribution is peaked at a finite extension independent of the size of the universe. Figs.~\ref{fig2}-\ref{fig3} pertain to space slices. Analogous results for time slices are summarized in Fig.~\ref{fig4}. In space slices of the lattice universe, one observes a transition from a percolating to a non-percolating phase at the Yang-Mills deconfinement temperature. Namely, in space slices, the vortex material distribution is strongly peaked at the maximal possible extension as long as the temperature remains below $T_C $; when the temperature rises above $T_C $, however, the distribution becomes concentrated at short lengths. The behavior near the deconfinement temperature $T_C $ displayed in Figs.~\ref{fig2}-\ref{fig3} deserves more detailed discussion. While the contents of the bin of maximal extension fall sharply between $T=0.8T_C $ and $T=1.1T_C $, a residual one quarter of vortex material remains concentrated in loops of maximal extension at the temperature identified as $T=1.1T_C $. This is too large a proportion to let pass by without further consideration. The authors have repeated the measurement at $T=1.1T_C $ on a larger, $16^3 \times 3$ lattice, and did not find a depletion of the bin of maximal extension. On the other hand, one should be aware that there is a considerable uncertainty, of the order of $10\% $, in the overall physical scale in these lattice experiments, affecting in particular the identification of the deconfinement temperature $T_C $ itself. These uncertainties were already mentioned in section \ref{centdo} and are discussed in detail in \cite{tempv}. At the present level of accuracy, $T=1.1T_C $ cannot be considered significantly separated from $T_C $; the authors cannot state with confidence that the measurement formally identified with a temperature $T=1.1T_C $ must unambiguously be associated with the deconfined phase. Note that also in standard string tension measurements via the Polyakov loop correlator, one does not attain a sharper signal of the deconfinement transition if one uses comparable lattices and statistics. Indeed, in \cite{tempv}, the authors still extracted a string tension of about $10\% $ of the zero-temperature value at the temperature formally identified as $T=1.1T_C $. In balance, the authors would argue that the percolation transition in space slices does occur together with the deconfining transition, both in view of the strong heuristic arguments connecting the two phenomena in the vortex picture, and in view of the sharp change in the vortex material distributions between $T=0.8T_C $ and $T=1.1T_C $. The latter sharp change suggests that the vortex material distributions can in practice be used as an alternative order parameter for the deconfinement transition. When the vortices rearrange at the transition temperature to form a non-percolating phase, intersection points of vortices with planes containing Polyakov loop correlators occur in pairs less than a maximal distance apart. This leads to a perimeter law for the Polyakov loop correlator, implying deconfinement. Consider now by contrast the vortex material distributions obtained in time slices. According to Fig.~\ref{fig4}, these distributions are strongly peaked at the maximal possible extension at all temperatures, even above the deconfinement transition. Thus, vortex line clusters in time slices always percolate; there is no marked change in their properties as the temperature crosses $T_C $. \begin{figure}[h] \centerline{ \epsfysize=9cm \epsffile{tcl.ps} } \caption{Vortex material distributions analogous to Fig.~\ref{fig2}, but taken from time slices of the $12^3 \times N_t $ lattice universe, again at inverse coupling $\beta =2.4$. Bins corresponding to three different temperatures are shown simultaneously at each cluster extension; namely, the case $N_t = 3$, which is identified with $T=1.85T_C $, the case $N_t = 5$, which is identified with $T=1.1T_C $, and the case $N_t = 7$, which is identified with $T=0.8T_C $.} \label{fig4} \end{figure} Note that this entails no consequences for the behavior of the Polyakov loop correlator, since Polyakov loops do not lie within time slices. However, the persistence of vortex percolation into the deconfined phase when time slices are considered represents one way of understanding the persistence of a spatial string tension above $T_C $. Given percolation, it seems plausible that intersection points of vortices with spatial Wilson loops continue to occur sufficiently randomly to generate an area law. There is another, complementary, way of understanding the spatial string tension which will be discussed in detail in the concluding section. Note furthermore that Figs.~\ref{fig2}-\ref{fig4} taken together imply that the vortices, regarded as two-dimensional surfaces in four-dimensional space-time, percolate both in the confined and the deconfined phases; this was also observed in \cite{bertl}. Only by considering a space slice does one filter out the percolation transition in the topology of the vortex configurations. It should be emphasized that the percolation of the two-dimensional vortex surfaces in four-dimensional space-time in the deconfined phase does not negate the heuristic picture of deconfinement put forward above. Given that vortex line clusters in space slices cease to percolate in the deconfined phase, intersection points of vortices with planes extending in one space and the time direction necessarily come in pairs less than a maximal distance apart, regardless of whether the different vortex line clusters do ultimately connect if one follows their world sheets into the additional spatial dimension. It is this pair correlation of the intersection points which induces the deconfinement transition. \subsection{Winding vortices in the deconfined phase} In order to gain a more detailed picture of the deconfined regime, it is useful to carry out the following analysis. Consider again a space slice of the lattice universe, in which vortex line clusters are short in the deconfined regime. Consider in particular lattices of time extension $N_t a$ with odd $N_t $, where $a$ is the lattice spacing; in the following numerical experiment, $N_t =3$. On such a lattice, measure vortex material distributions akin to the ones described in the previous section, with one slight modification; namely, define the bins of the histograms not by cluster extension, but simply by the number of dual lattice links contained in the clusters. It turns out that, in the deconfined phase, specifically at $T=1.85T_C $, roughly $55\% $ of the vortex material is concentrated in clusters made up of an odd number of links, cf. Fig.~\ref{figb}. On a lattice with $N_t =3$, these are necessarily vortex loops which wind around the lattice in (Euclidean) time direction by virtue of the periodic boundary conditions, where the loops containing an odd number of links larger than $3$ exhibit residual transverse fluctuations in the spatial directions, as also visualized in Fig.~\ref{phases} further below. \begin{figure}[h] \centerline{ \epsfysize=9cm \epsffile{wcl.ps} } \caption{Vortex material distributions in space slices of the lattice universe as a function of total vortex line length contained in the clusters. On the $12^3 \times N_t =3 $ lattice used, clusters with a length of an odd number of lattice spacings necessarily wind around the lattice in the Euclidean time direction. The inverse coupling in this measurement was again set to the value $\beta =2.4$, implying that this measurement is associated with a temperature of $T=1.85T_C $. There is a residual, but insignificant, proportion of vortex clusters containing more than $20$ dual lattice links not displayed in the plot.} \label{figb} \end{figure} One thus obtains a quite specific characterization of the short vortices appearing in the deconfined regime. This phase can evidently be visualized largely in terms of short winding vortex loops with residual transverse fluctuations if one considers a space slice of the lattice universe, cf. Fig.~\ref{phases}. Note that this picture also explains the partial vortex polarization observed in density measurements \cite{tempv}. \begin{figure}[h] \centerline{ \epsfysize=7cm \epsffile{phas.ps} \vspace{1cm} } \caption{Visualization of typical vortex configurations determining the long-range physics of the confined and deconfined phases of Yang-Mills theory. Note that this is not a depiction of particular configurations found in lattice experiments; rather, it is the authors' interpretation of the measurements shown in Figs.~\ref{fig2}-\ref{figb} in terms of typical configurations dominating the Yang-Mills functional integral in the confined and deconfined phases. Shown are space slices of the lattice universe obtained by holding the $z$-coordinate fixed. Slicing the two-dimensional vortex surfaces present in four space-time dimensions yields one-dimensional loop configurations such as depicted.} \label{phases} \end{figure} \section{Discussion and Outlook} On the basis of the measurements shown in the preceding sections, a detailed description of the confined and deconfined phases of Yang-Mills theory in terms of center vortices emerges. The typical vortex configurations present in the two phases are visualized in Fig.~\ref{phases}. This picture allows an intuitive understanding of the phenomenon of confinement as well as the characteristics of the transition to the deconfined phase. In the confined phase, vortex line clusters in space slices of the lattice universe percolate. This allows intersection points of vortices with planes containing Polyakov loop correlators to occur sufficiently randomly to generate an area law. By contrast, in the deconfined phase, typical vortex configurations in space slices of the lattice universe are characterized by short vortex loops, to a large part winding in the (Euclidean) time direction. This causes intersection points of vortices with planes containing Polyakov loop correlators to occur in pairs less than a maximal distance apart, leading to a perimeter law. Simple analytical model arguments clarifying the emergence of this qualitative difference were presented in the introduction. The deconfinement phase transition in the vortex picture can thus be understood as a transition from a percolating to a non-percolating phase. It should be emphasized that the percolation properties of vortices focused on in the present work are more stringently related to confinement than the polarization properties reported in \cite{tempv}. There is a priori no direct logical connection between the observed partial vortex polarization by itself and deconfinement. On the one hand, even in presence of a significant polarization, confinement would persist as long as the vortex loops retain an arbitrarily large length, namely by winding sufficiently often around the (Euclidean) time direction before closing. On the other hand, even in an ensemble with no polarization, deconfinement will occur if the vortices are organized into many small isolated clusters. Thus, vortex polarization should be viewed more as an accompanying effect than the direct cause of deconfinement. Of course, a correlation between the absence of percolation in space slices of the lattice universe and vortex polarization is not surprising. If fluctuations of vortex loops in the space directions are curtailed, e.g. due to a phase containing many short vortices winding in the time direction becoming favored (more about this below), then clearly the connectivity of vortex clusters in the space direction is reduced and they may cease to percolate. In this sense, polarization indirectly can facilitate deconfinement. However, the percolation concept is related much more directly and with much less ambiguity to the question of confinement. Ultimately, this is a consequence of a point already made in the introduction in connection with the heuristic models discussed there. Since the Wilson loop should be independent of the choice of area which one may regard it to span, it is conceptually sounder not to consider densities occuring on such areas, but the global topology of the vortices such as their linking number with the Wilson loop. The likelihood of a particular linking number occuring is strongly influenced by the connectivity of the vortex networks. Correspondingly, there is a clear signal of the phase transition in the vortex material distributions displayed in Figs.~\ref{fig2}-\ref{fig3}; these quantities can be used as alternative order parameters for the transition. By contrast, the vortex densities seem to behave smoothly across the deconfinement phase transition \cite{tempv}. Turning to the spatial string tension, there are two complementary ways to qualitatively account for its persistence in the deconfined phase of Yang-Mills theory. One was already mentioned in section \ref{secsp}. If one considers a time slice of the lattice universe, the associated vortex line configurations display no marked change of their clustering properties across the deconfinement transition. Even in the deconfined phase, vortex loops in time slices percolate. In view of this, it seems plausible that intersection points of vortices with spatial Wilson loops continue to occur sufficiently randomly to generate an area law. It should be noted, however, that this percolation is qualitatively different from the one observed in the confined phase in that it only occurs in the three space dimensions, whereas the configurations are relatively weakly varying in the Euclidean time direction. In other words, in the deconfined phase, one finds a dimensionally reduced percolation phenomenon only visible either in the full four space-time dimensions or in time slices thereof. On the other hand, if one considers a space slice of the lattice universe, the deconfined phase is characterized to a large part by short vortex loops winding in the time direction, cf. Fig.~\ref{phases}. However, in this topological setup, such short vortices can pierce the area spanned by a large spatial Wilson loop an odd number of times, even far from its perimeter. This should be contrasted with the picture one obtains for the Polyakov loop correlator. There, shortness of vortices implies that their intersection points with the plane containing the Polyakov loop correlator occur in pairs less than a maximal distance apart. This leads to a perimeter law behavior of the Polyakov loop correlator, i.e. deconfinement. For spatial Wilson loops, this mechanism is inoperative due to the different topological setting. On the contrary, in view of Fig.~\ref{phases}, if one assumes the locations of the various winding vortices to be uncorrelated, one obtains precisely the heuristic model of the introduction, in which vortex intersection points are distributed at random on the plane containing the spatial Wilson loop, leading to an area law. Finite length vortex loops thus do not contradict the existence of a {\em spatial} string tension. Of course, there is no reason to expect the locations of the winding vortices to be completely uncorrelated in the high-temperature Yang-Mills ensemble. In fact, comparing the values for the spatial string tension $\kappa_{s} $ from \cite{karsch} and the relevant density $\rho_{s} $ of vortex intersection points on planes extending in two spatial directions \cite{tempv}, the ratio $\kappa_{s} /\rho_{s} $ reaches values $\kappa_{s} /\rho_{s} \approx 3 $ at $T\approx 2T_C $. This should be contrasted with the value $\kappa =2\rho $ obtained in the model of random intersection points discussed in the introduction. If one further takes into account that a sizeable part of $\rho_{s} $ is still furnished by non-winding vortex loops, cf. Fig.~\ref{figb}, then one should actually use the density $\rho_{s}^{\prime } < \rho_{s} $ corresponding to winding vortices only in the above consideration. This yields an even larger ratio $\kappa_{s} /\rho_{s}^{\prime } $. Therefore, the winding vortices in the deconfined phase seem to be subject to sizeable correlations. Both of the above complementary mechanisms generating the spatial string tension in the deconfined phase are qualitatively distinct from the mechanism of confinement below $T_C $. In the space-slice picture, this is obvious; a new class of configurations, namely short vortex loops winding in the Euclidean time direction, induces the spatial string tension. However, as already indicated further above, also in the time-slice picture, the observed percolation is qualitatively different from the one in the confined phase in that it is dimensionally reduced. This qualitatively different origin of the spatial string tension may provide a natural explanation for the novel behavior detected in section \ref{centdo} for spatial Creutz ratios at temperatures well inside the deconfined regime; namely, their rise as a function of the size of the Wilson loops from which they are extracted (as opposed to the precocious scaling observed at lower temperatures). However, the detailed connection between the abovementioned modified dynamics in the deconfined phase and the signal seen in the measurements of spatial Creutz ratios remains unclear. While the relevant characteristics of the vortex configurations in the different regimes were described in detail in this work, the present understanding of the underlying dynamics in the vortex picture is still tenuous. There are, however, indications that the deconfining percolation transition can be understood in terms of simple entropy considerations. Increasing the temperature implies shortening the (Euclidean) time direction of the (lattice) universe. This means that the number of possible percolating vortex configurations decreases simply due to the reduction in space-time volume\footnote{For example, a vortex surface extending into two space directions has a greatly reduced freedom of transverse fluctuation into the time direction. Note that if one thinks of such a fluctuating, fuzzy thin vortex surface in terms of a thick envelope, this amounts to stating that the thick vortex extending into two space directions simply does not fit into the space-time manifold anymore. To a certain extent, the difference between these two (thin and thick vortex) pictures is semantic. To state that a thick vortex does not fit into the space-time manifold perpendicular to the time direction amounts to nothing but the statement that the number of possible configurations of this type has been reduced (to zero).}. At the same time as the number of possible percolating vortex clusters is reduced, the number of available short vortex configurations is enhanced by the emergence of a new class of short vortices at finite temperature, namely the vortices winding in time direction. In view of this, it seems plausible that a transition to a non-percolating phase is facilitated as temperature is raised. There are two pieces of evidence supporting this explanation, one of which was already given above. Namely, the deconfined phase indeed contains a large proportion of short winding vortices, cf. Fig.~\ref{figb}. More than half of the vortex material is transferred to the newly available class of short winding vortices in the deconfined phase. The second piece of evidence is related to the behavior of stiff random surfaces in four space-time dimensions; some of the authors plan to report on their Monte Carlo investigation of these objects in an upcoming publication. The model assumes that the vortices are random surfaces associated with a certain action cost per unit area and a penalty for curvature of the vortex surface. By construction, evaluating the partition function of this model simply corresponds to counting the available vortex configurations under certain constraints imposed by the action; namely, the action cost per surface area effectively imposes a certain mean density of vortices, while the curvature penalty imposes an ultraviolet cutoff on the fluctuations of the vortex surfaces. Beyond this, no further dynamical information enters. It turns out that already this simple model generates a percolation phase transition analogous to the one observed here for the center vortices of Yang-Mills theory. This suggests that the deconfining percolation transition of center-projected Yang-Mills theory can be understood in similarly simple terms, without any need for detailed assumptions about the form of the full center vortex effective action. \section{Acknowledgements} Discussions with F.~Karsch and H.~Satz are gratefully acknowledged. K.L. also acknowledges the friendly hospitality of the members of the KIAS, Korea, where a part of the numerical computations was carried out.
\section{Introduction} The standard description of the spacetime geometry in General Relativity uses the metric tensor $g_{\mu \nu }$ as the fundamental field. In hamiltonian form, the action is a functional of the spatial metric $h_{ij}$ and its canonical momentum $\pi ^{ij}$, as well as four Lagrange multipliers associated with spatial reparametrizations, $N^{i}$ (shifts) and normal deformations, $N^{\bot \text{ }}$(lapse) \cite{ADM,Claudio}\footnote{% In this analysis the torsion tensor is assumed to vanish identically.}. The Einstein-Hilbert action can also be written in terms of the spin connection $\omega _{\mu }^{ab}$, the tetrad field $e_{\nu }^{a}$ and their exterior derivatives (first order formalism)\cite{zumino,regge}. In this form, the hamiltonian construction needed to identify the dynamical degrees of freedom is not straightforward. The torsion tensor does not vanish identically, but only as a consequence of the classical equations. Thus, it is not legitimate to eliminate the spin connection for the vierbein and one is left with a theory for 40 independent fields $\omega _{i}^{ab}$ (18), and $e_{i}^{a}$ (12). Also, many of the fields do not have time derivatives: out of the 40, only 12 have time derivatives in the lagrangian. This gives rise to a number of first and second class constraints. \subsection{The two frames} An additional difficulty is that there are two natural choices of coordinates and momenta, which have radically different phase space and constraint structure. Thus, two options arise, depending on whether the lagrangian involves only time derivatives of the vierbein ($e$-frame), or the spin connection ($\omega$-frame). These two choices are related by a canonical transformation (they differ by a total derivative) and therefore should be classically identical in content\footnote{% It is extremely difficult to establish the quantum mechanical equivalence between canonically related formulations of a theory, and the equivalence may not even exist. It has been shown that quantum mechanics could be formulated in a way that is invariant under the simpler class of point canonical transformations \cite{JL}, but a similar proof for general canonical transformations is not yet known.}. However, as the corresponding phase spaces are so radically different, proving the equivalence even at the classical level is non trivial. To make the discussion more concrete, let us recall some facts about the first order formulation of gravity. The first order Lagrangian is the Einstein-Hilbert four-form (wedge product of forms is understood) \begin{equation} L_{E-H}=\epsilon _{abcd}R^{ab}e^{c}e^{d}+dB, \label{EH} \end{equation} where $R^{ab}=d\omega ^{ab}+\omega _{c}^{a}\omega ^{cb}$ is the curvature two-form, $\omega $ is the spin connection one-form, $e$ is the vierbein one-form and $dB$ stands for some arbitrary boundary term. Different choices of $B(e,\omega )$ give rise to different choices of canonical coordinates (frames). Two natural choices are: \subsubsection{\bf The e-frame \newline } In the hamiltonian analysis of this action in first order form \cite{BC} the spin connection splits in two pieces. One of them corresponds to the canonical momentum of the tetrad field, and the other corresponds to a set of auxiliary variables that can be eliminated from their own equations of motion in terms of the tetrad. The resulting hamiltonian is a functional of the tetrad and its canonical momentum only, and the spin connection drops out. In this way the usual vierbein formulation of gravity is obtained \cite {marc}. \subsubsection{{\bf The }$\omega ${\bf -frame\newline }} Alternatively, one can start from the Einstein-Hilbert action, eliminate the tetrad field and build a hamiltonian action that depends on the spin connection and its canonical momentum only \cite{Peldan}. A preliminary discussion of the equivalence between the $\omega $ and $e$-frames was presented in \cite{Contreras}. In this letter we want address some points of the analysis in the $\omega $-frame. \subsubsection{\bf The Ashtekar approach\newline } The alternative approach to canonical gravity proposed by Ashtekar \cite {Ashtekar} in the past decade is yet another canonically equivalent description of General Relativity. The Ashtekar frame is obtained through a complex canonical transformation from the $e$-frame \cite{HNS}. It has been often discussed whether Ashtekar's theory is quantum mechanically equivalent to standard metric gravity and the answer still seems uncertain and possibly irrelevant. As we show here, the $\omega $ and $e$-frames are not only equivalent classically through a real canonical transformation but, if there were a quantum description for either one, it would be equivalent to the quantum description for the other. \section{First order formalism (in the $\omega$ frame)} As shown in \cite{Peldan,Contreras}, dropping the boundary term in (\ref{EH}% ), the first order action for gravity in 3+1 dimensions can also be written as\footnote{% Our conventions are that $\epsilon ^{ijk}=\pm 1,0$ is a tensor density of weight 1 (i.e., it transforms like a tensor of third rank times $\sqrt{g}$). Hence, $\epsilon _{ijk}\equiv g_{il}g_{jm}g_{kn}\epsilon ^{lmn}$ is also a tensor density of weight 1, but it takes values $\pm g,0$.} \begin{equation} I[w,e]=\int (\dot{\omega}_{k}^{ab}\epsilon _{abcd}\epsilon ^{ijk}e_{i}^{c}e_{j}^{d}+\omega _{0}^{ab}J_{ab}+e_{0}^{a}P_{a}), \label{first} \end{equation} where $J_{ab}=\epsilon _{abcd}\epsilon ^{ijk}T_{ij}^{c}e_{k}^{d}=D_{k}(\epsilon _{abcd}\epsilon ^{ijk}e_{i}^{c}e_{j}^{d})$, $P_{a}=(\epsilon _{abcd}\epsilon ^{ijk}R_{ij}^{bc}e_{k}^{d})$. If $e_{0}^{a}$ is descomposed as $% e_{0}^{a}=N\eta ^{a}+N^{i}e_{i}^{a}$, where $\eta ^{a}$ is the normal to espacelike surfaces, $\eta _{a}e_{i}^{a}=0$, the action can be written in terms of the $\omega _{i}^{ab}$ and its canonically conjugate momentum $% P_{ab}^{k}=\epsilon _{abcd}\epsilon ^{ijk}e_{i}^{c}e_{j}^{d}$, in the form \begin{equation} I=\int (\dot{\omega}_{k}^{ab}P_{ab}^{k}-\omega _{0}^{ab}J_{ab}+NH_{\perp }+N^{i}H_{i}+\mu _{ij}\phi ^{ij}), \end{equation} where \begin{equation} H_{\perp }=g^{-1/2}P_{ac}^{i}P_{b}^{cj}R_{ij}^{ab}, \end{equation} \begin{equation} H_{i}=P_{ab}^{j}R_{ij}^{ab}-\omega _{i}^{ab}J_{ab}, \end{equation} \begin{equation} J_{ab}=D_{i}P_{ab}^{i}, \end{equation} and \begin{equation} \phi ^{ij}=\epsilon ^{abcd}P_{ab}^{i}P_{cd}^{j}. \end{equation} Here $\omega _{0}^{ab}$, $N$, and $N^{i}$ are Lagrange multipliers corresponding to the constraints $J_{ab}=H_{\perp }=H_{i}=0$, and $% g=det(g_{ij})$, $g_{ij}\equiv e_{i}^{a}e_{aj}$. The presence of the constraint $\phi^{ij}=0$ deserves some discussion. The substitution of $P_{a b}^k$ for $\epsilon^{i j k} \epsilon_{a b c d} e^c_i e^d_j $ conceals the fact that there are only 12 independent fields ($e^a_i$% ) and not 18 ($P_{a b}^i$) in the phase space. The elimination of the 6 spurious fields is enforced by the 6 conditions $\phi^{i j} = 0$. The Jacobian of the transformation $e^a_i \rightarrow P_{a b}^i$ is $\Omega_{a b \ c}^{i \ \ j} = 2 \epsilon_{a b c d} \epsilon^{i j k} e^d_k$, which has maximun rank (twelve) on configurations for which the local orthonormal frames $e^a_i$ are generic, that is, they span a 3-dimensional volume (see below). Once the second class constraints have been eliminated, $H_i$ and $J_{a b}$ become the generators of spatial diffemorphism and local rotations, respectively. Preservation in time of the constraint $\phi ^{ij}=0$ implies a new constraint \begin{equation} \chi ^{kl}\delta (x,y)=\{\phi ^{kl}(x),H_{\perp }(y)\}=g^{-1/2}D_{i}(P_{ec}^{(k})P_{ab}^{l)}P_{f}^{ci}\epsilon ^{abfe}\delta (x,y), \label{chi} \end{equation} where the parentheses indicate symmetrization in $k$, $l$. Preservation of $% \chi ^{kl}=0$ in turn, implies \begin{equation} N\{H_{\perp },\chi ^{kl}\}+\mu _{mn}\{\phi ^{mn},\chi ^{kl}\}=0. \label{chidot} \end{equation} These are equation for the Lagrange multipliers, which can be solved for $% \mu $ because the constraints $\phi ^{mn}=\chi ^{kl}=0$ obey a second class algebra, \begin{equation} \begin{array}{ll} \{\phi ^{mn}(x),\phi ^{kl}(y)\} & =0\nonumber \\ \{\chi ^{mn}(x),\chi ^{kl}(y)\} & \neq 0\nonumber \\ \{\chi ^{ij}(x),\phi ^{mn}(y)\} & =g^{-1/2}B^{ijmn}(x,y) \\ & =g^{-1/2}(G^{ijmn}(\tilde{g}(x))+G^{ijmn}(\phi (x)))\delta (x-y).\label% {second} \end{array} \end{equation} where $G^{ijmn}(A)$ is the inverse supermetric for a symmetric matrix $% A^{ij} $ \footnote{% Here, the metric is not defined yet, $\tilde{g}^{ij}$ is just a shorthand for $P_{ab}^{i}P^{abj}$ which will eventually be related to the canonical metric through $\tilde{g}^{ij}=-8gg^{ij}$.} \begin{equation} G^{ijkl}(A)=2A^{ij}A^{kl}-A^{il}A^{kj}-A^{ik}A^{lj}. \end{equation} \newline The precise form of $\{\chi ,\chi \}$ is not essential as we will see below. The matrix $B^{ijmn}(x,y)$ has a formal inverse $B_{ijmn}$ given by the series \begin{equation} B=G(g)-G(g)G^{-1}(\phi )G(g)+G(g)G^{-1}(\phi )G(g)G^{-1}(\phi )G(g)+... \end{equation} where we have defined $G\sim G_{ijkl}$, $B\sim B_{ijkl}$, $G^{-1}\sim G^{ijkl}$, etc. Obviously $B_{ijmn}$ coincides with the supermetric $G_{ijmn}$ on the constraint surface $\phi =0$, $\chi =0$. In this way we can solve (\ref {chidot}) for $\mu _{ij}$, \begin{equation} \mu _{ij}=\frac{1}{2}B_{ijmn}N\{H_{\perp },\chi ^{mn}\}. \end{equation} Thus no new constraints appear from the preservation in time of $\chi $. The initial $H_{\perp }$ has nonvanishing Poisson brackets with $\phi $ or $\chi $, but the modified one \begin{equation} \tilde{H}_{\perp }=H_{\perp }-\frac{1}{N}\mu _{ij}\phi ^{ij}, \end{equation} does. \section{Dirac brackets} The second class constraints can be eliminated through Dirac bracket \cite {Dirac,HT} defined by \begin{equation} \{ U, V \}^{*} = \{ U, V \} - \{ U, \varphi^{\alpha} \} C_{\alpha \beta} \{ \varphi^{\beta}, V \} , \label{cordi} \end{equation} where $C_{\alpha \beta}$ is the inverse of the Dirac matrix $C^{\alpha \beta} = \{ \varphi^{\alpha} , \varphi^{\beta} \}$, where $\varphi^{\alpha}$ denote generic second class constraints. In our case, the Dirac matrix \begin{equation} C^{\alpha \beta} (x,y) = \left( \begin{array}{cc} \{ \chi^{i j}(x) , \chi^{m n}(y) \} & \{ \chi^{i j}(x) , \phi^{m n}(y) \} \\ \{ \phi^{i j}(x) , \chi^{m n}(y) \} & \{ \phi^{i j}(x) , \phi^{m n}(y) \} \end{array} \right) , \label{dmx} \end{equation} has the form $\left( \begin{array}{cc} A & B^{-1} \\ - B^{-1} & 0 \end{array} \right) $, because $\{ \phi^{i j}, \phi^{m n} \} = 0 $. \newline Its inverse takes the form $C^{-1} = \left( \begin{array}{cc} 0 & - B \\ B & B A B \end{array} \right) $, or \begin{equation} C^{-1}_{\alpha \beta} = \left( \begin{array}{cc} 0 & - B_{i j m n} \\ B_{i j m n} & B_{i j p q} \{ \chi^{p q} , \chi^{k l} \} B_{k l m n} \end{array} \right) . \end{equation} The Dirac bracket for two arbitry functionals U, V is given by: \begin{equation} \begin{array}{ll} \{ U , V \}^{*} = & \{ U , V \} - \{ U , \chi \} B \{ \phi, V \} \\ & - \{ U , \phi \} B \{ \chi, V \} - \{ U , \phi \} B \{ \phi, \phi \} B \{ \phi, V \} , \end{array} \label{abd} \end{equation} where sum and integration over discrete and continuous indices is assumed. \newline It can be shown that when U and V belong to the set $\{\tilde{H}_{\perp} , H_i , J_{a b} \} $, the second term of the right hand side of (\ref{abd}) vanishes on the constraint surface $\phi = 0$, $\chi = 0$. In particular, direct subtitution in (\ref{abd}) yields \begin{equation} \{ \tilde{H}_{\perp}, \tilde{H}_{\perp}\}^* = \{H_{\perp}, H_{\perp}\}^* = \{H_{\perp}, H_{\perp}\}. \end{equation} and using the results of \cite{Peldan}, we finally have \begin{equation} \{ H_{\perp}, H_{\perp} \}^{*} \sim g^{i j} H_j \partial_i \delta(x,y) . \end{equation} In the same way, the complete Dirac algebra can be shown to be given by \begin{equation} \{ H_{\perp}[N] , H_{\perp}[M] \}^{*} = \int [ (\partial_i N) M - (\partial_i M) N ] g^{i j}(P) H_j , \label{h-h} \end{equation} \begin{equation} \{ H_{\perp} [N] , H_i [M^{i}] \}^{*} = \int ( M^{i} \partial_{i} N - N \partial_{i} M^{i} ) H_{\perp} , \end{equation} \begin{equation} \{ H_i [M^{i}] , H_j [M^{j}] \}^{*} = \int (N^{l} \partial_{l} M^{m} - M^{l} \partial_{l} N^{m}) H_{m} , \end{equation} \begin{equation} \{ J[N^{a b}] , J[M^{c d}] \}^{*} = \int J[(M \times N)^{a b}] , \end{equation} \begin{equation} \{ J[N^{a b}], H[M] \}^{*} = 0 , \end{equation} \begin{equation} \{ J[N^{a b}], H[M^{i}] \}^{*} = 0 . \label{j-h} \end{equation} Thus, the Dirac algebra reduces to a direct sum of the usual algebra of spacetime diffeomorphism plus tangent space rotations. Note that when $P_{a b}^k$ is replaced by its expresion in terms of the tetrad, the $\phi^{i j}$ constraints vanish identically, but the secondary constraints $\chi^{i j}$ do not. In the vierbein frame \cite{BC} it can also be shown that prior to eliminating the auxiliary variables, apart from $J_{a b}, H_{\perp}, H_{i}$ the constraints \begin{equation} \gamma^{i j} = E_{a}^{(i} \ \epsilon^{m n j)} \ T^{a}_{m n} = 0 , \label{tilphi} \end{equation} are found, where $T^{a}_{i j}$ are the spatial components of the torsion tensor, and $E_a^i \equiv e_{a j} \ g^{i j}$. Equation (\ref{tilphi}) is one of the field equations, from which the auxiliary variables $\lambda^{i j}$ can be eliminated. Replacing $P_{a b}^i = \epsilon^{ijk} \epsilon_{a b c d} e^c_j e^d_k $ in the definition (\ref{chi}), $\chi^{i j}$ can be identified with $\gamma^{i j}$ in the e-frame. The algebra (\ref{h-h}--\ref{j-h}) is the same as the one found in the vielbein-frame once the contraint $\gamma^{ij}$ is strongly set equal to zero. The two frames can be compared and contrasted in the following table: \newline \newline \newline \begin{tabular}{|c|c|c|} \hline \ \ \ & e-frame & $\omega$-frame \\ \hline & & \\ dynamical & $e^a_i, \pi_a^i = \epsilon^{ijk} \epsilon_{a b c d} e^b_j \omega^{c d}_k $ & $\omega^{a b}_i, P_{a b}^i = \epsilon^{ijk} \epsilon_{a b c d} e^c_j e^d_k $ \\ variables & & \\ (q,p) & (12) , (12) & (18) , (18) \\ & & \\ \hline First class & & \\ constraint & $H_{\perp}, H_{i}, J_{a b}$ & $H_{\perp}, H_{i}, J_{a b}$ \\ & & \\ \hline second class & & \\ constraint & --------- & $\chi^{ij}, \phi^{kl}$ \\ & & \\ \hline prop. & & \\ degrees of & 12 - 4 - 6 = 2 & 18 - 10 - $\frac{1}{2}$ 12 = 2 \\ freedom & & \\ & & \\ \hline \end{tabular} \newline \newline (Here $\gamma^{ij}$ has been eliminated in the $e$-frame). The number of propagating degrees of freedom is $g = c - f - \frac{1}{2}s $, where $c$ is the number of coordinates, $f$ the number of first class constraints and $s$ the number of second class constraints. \section{Path integral} We now consider the path integral for this system. As shown in \cite{HT}, the path integral for a system with second class constraints $\chi$, $\phi$ has a measure proportional to \begin{equation} \delta (\chi) \delta(\phi) \sqrt{det C^{\alpha \beta}} , \label{measure} \end{equation} where $C^{\alpha \beta}$ is the Dirac matrix. In our case, $C^{\alpha \beta}$ as given in (\ref{dmx}) and (\ref{second}), yields \begin{equation} \sqrt{det C^{\alpha \beta} } = det( G^{i j m n}(\tilde{g})+G^{i j m n}(\phi) ) . \end{equation} The delta functions in (\ref{measure}) restrict the integration to the constraint surface $\phi = 0$, $\chi = 0$, so path integral reads \begin{equation} Z = \int [DP_{a b}^{k}] [D\omega^{a b}_{k}] det(G^{ijmn}(\tilde{g})) \delta(\phi^{ij}) \delta(\chi^{mn}) det M_{\alpha \beta} \delta(H_{\perp}) \delta(H_{i}) \delta(J_{a b}) \ exp \frac{i}{\hbar} S \label{zeta} \end{equation} with \begin{equation} S = \int \dot{\omega}^{a b}_{k} P_{a b}^{k} , \end{equation} and $M_{\alpha \beta}$ is the matrix of Poisson brackets \begin{equation} M_{\alpha \beta} = \{ F_{\alpha} , \varphi_{\beta} \}^* , \end{equation} where $F_{\alpha}$ are gauge condition for the first class constraint set $% \varphi_{\beta} = \{ H_{\perp}, H_i , J_{a b} \}$. \section{{\bf The }$\omega ${\bf -}$e${\bf \ transformation}} Consider now the following transformation, wich maps the 18 coordinates $% \omega _{i}^{ab}$ and their 18 canonically conjugate momenta $P_{ab}^{i}$ into 12 $e_{i}^{a}$'s, 12 $\pi _{a}^{i}$'s, 6 auxiliary variables $\lambda _{mn}$ and 6 $\rho ^{mn}$ ( $\lambda _{mn}$ and $\rho ^{mn}$ are symmetric and m,n take the values 1,2,3.) \begin{equation} \omega _{k}^{ab}=\Theta _{k\ \ j}^{ab\ c}\ \pi _{c}^{j}+U_{k}^{ab\ \ mn}\ \lambda _{mn} \label{trans1} \end{equation} \begin{equation} P_{ab}^{k}=\frac{1}{2}\Omega _{ab\ c}^{k\ \ j}\ e_{j}^{c}+V_{ab\ \ mn}^{k}\ \rho ^{mn}. \label{trans2} \end{equation} Here $\Theta $ and $\Omega $ are rectangular matrices, \begin{equation} \Theta _{i\ \ j}^{ab\ c}=\frac{1}{8\sqrt{g}}[e_{i}^{[a}\eta ^{b]}e_{j}^{c}-e_{i}^{[a}e_{j}^{b]}\eta ^{c}-2e_{j}^{[a}\eta ^{b]}e_{i}^{c}], \end{equation} \begin{equation} \Omega _{ab\ c}^{i\ \ j}=2\epsilon _{abcd}\ e_{k}^{d}\ \epsilon ^{ijk}, \end{equation} where the square brackets indicate antisymmetrization. $U$ and $V$ are null vectors for $\Omega $ and $\Theta $, given by \begin{equation} U_{k}^{ab\ \ mn}=\frac{1}{2}\delta _{i}^{(m}\ \epsilon ^{n)kl}\ e_{k}^{a}\ e_{l}^{b}, \end{equation} \begin{equation} V_{ab\ \ mn}^{k}=\frac{1}{g}\ E_{a}^{r}\ E_{b}^{s}\epsilon _{rs(m}\ \delta _{n)}^{i}. \end{equation} These objects satisfy the following relations \begin{equation} \Omega _{ab\ c}^{k\ \ i}\ \Theta _{k\ \ j}^{ab\ d}=\delta _{d}^{c}\delta _{j}^{i}, \label{orto1} \end{equation} \begin{equation} \Omega _{ab\ c}^{k\ \ j}\ U_{k}^{ab\ \ mn}=0, \end{equation} \begin{equation} \Theta _{k\ \ j}^{ab\ c}\ V_{ab\ \ mn}^{k}=0, \end{equation} \begin{equation} U_{k}^{ab\ \ mn}\ V_{ab\ \ pq}^{k}=\delta _{(pq)}^{(mn)}. \label{orto4} \end{equation} One can think of $\Theta $ and $\Omega $ as a collection of twelve vectors --labeled by the indices $(_{i}^{a})$ and $(_{a}^{i})$ respectively--, in an 18-dimensional vector space with components $(_{j}^{ab})$, and $(_{ab}^{j})$% , respectively. By the same token, $U$ and $V$ are six vectors (labeled by the index $(mn)$) in an 18-dimensional vector space. In this sense the properties (\ref{orto1},...\ref{orto4}) are nothing but orthogonality relations among the vectors $\Theta$, $\Omega $, $U$ and $V$. These relations imply the following completeness relation \begin{equation} \Theta^{ab \ e}_{i \ \ l} \ \Omega_{cd \ e}^{j \ \ l} + U^{ab \ \ mn}_{i} \ V_{cd \ \ mn}^{j} = \delta^{[a b]}_{[c d]} \delta^i_j , \label{completeness} \end{equation} which will be used in what follows. In this way, the 18 vectors $\Theta^{ab \ c}_{i \ \ j}$, $U^{ab \ \ mn}_{i}$ are a basis of the space of contravariant vectors $L^{ab}_i$, while $\Omega_{ab \ c}^{i \ \ j}$, $V_{ab \ \ mn}^{j}$ are a basis for the dual (covariant vectors, $L_{ab}^i$). Thus, the field transformations (\ref{trans1}), (\ref{trans2}) correspond to the expansions of $\omega^{ab}_i$ and $P_{ab}^i$ in the contravariant and covariant bases, respectively. As shown in the appendix, using (\ref{trans1}), (\ref{trans2}) the path integral (\ref{zeta}) can now be written in terms of the coordinates of the $% e$-frame as \begin{equation} Z=\int [De_{i}^{a}][D\pi _{c}^{j}]\ detM_{\alpha \beta }\ \delta (H_{\perp })\delta (H_{i})\delta (J_{ab})\ exp\frac{i}{\hbar }S, \end{equation} which is the path integral one would write in the $e$-frame. This shows the equivalence between quantum theories one would obtain in the two frames. The different constraints $H_{\perp }$, $H_{i}$,and $J_{ab}$ can be written explicitly in terms of $e$-frame variables, as \begin{equation} \frac{1}{2}H_{\perp }=\eta ^{a}\partial _{i}\pi _{a}^{i}-\frac{1}{2}% E_{d}^{s}\partial _{[l}e_{s]}^{d}\eta ^{b}\pi _{b}^{l}-G_{\perp ij}^{ab}\pi _{a}^{i}\pi _{b}^{j}-g^{3/2}G^{mnpq}\lambda _{mn}^{0}\lambda _{pq}^{0}, \end{equation} \begin{equation} \begin{array}{ll} N^{m}H_{m}= & N^{m}[\frac{1}{2}(g^{-1}E_{d}^{s}\partial _{i}e_{k}^{d}\epsilon _{mls}\epsilon ^{ijk}e_{j}^{b}-E_{d}^{s}\partial _{[m}e_{s]}^{d}e_{l}^{b}+\eta _{d}\partial _{[m}e_{l]}^{d}\eta ^{b})\pi _{b}^{l} \\ & +e_{m}^{a}\partial _{i}\pi _{a}^{i}+G_{mij}^{ab}\pi _{a}^{i}\pi _{b}^{j}]+% \frac{1}{2}N^{(m}e_{i}^{a}e_{j}^{b}J_{ab}\epsilon ^{ijn)}\lambda _{mn}^{0} \\ & -N^{i}\omega _{i}^{ab}\ J_{ab}, \end{array} \end{equation} \begin{equation} J_{ab}=2\epsilon _{abcd}\frac{\partial e_{j}^{c}}{\partial x^{i}}% e_{k}^{d}\epsilon ^{ijk}-\frac{1}{2}(\pi _{a}^{i}e_{bi}-\pi _{b}^{i}e_{ai}), \end{equation} where \begin{equation} \lambda _{pq}^{0}=\frac{1}{2g}G_{pqmn}E_{a}^{(m}\partial _{i}e_{j}^{a}\epsilon ^{ijn)}, \label{lambda0} \end{equation} \begin{equation} G_{\perp ij}^{ab}=\frac{1}{16\sqrt{g}}[% e_{i}^{a}e_{j}^{b}-2e_{j}^{a}e_{i}^{b}-g_{ij}\eta ^{a}\eta ^{b}], \end{equation} and \begin{equation} G_{mij}^{ab}=\frac{1}{16\sqrt{g}}[g_{ij}\eta ^{a}e_{m}^{b}+2g_{im}(e_{j}^{a}\eta ^{b}-e_{j}^{b}\eta ^{a})]. \end{equation} Finally, the kinetic term $P_{ab}^{i}\dot{\omega}_{i}^{ab}$ in the action S reduces via the $e$-$\omega $ transformation to the usual $e$-frame kinetic term $\pi _{a}^{i}\dot{e}_{i}^{a}$. This completes the classical and quantum equivalence between the $\omega $ and $e$ frames. \begin{quotation} {\LARGE Acknowledgments} \end{quotation} The authors are grateful to M. Ba\~{n}ados, M. Henneaux and R. Troncoso for many enlightening discussions and helpful comments. This work was supported in part through grants 1960229, 1970151, 1980788 3960007, and 7960001 from FONDECYT, and grant 27-953/ZI-DICYT (USACH). The institutional support of FUERZA AEREA DE CHILE, I.\ Municipalidad de Las Condes, and a group of Chilean companies (AFP Provida, Business Design Associates, CGE, CODELCO, COPEC, Empresas CMPC, GENER\ S.A., Minera Collahuasi, Minera Escondida, NOVAGAS and XEROX-Chile) is also recognized.
\section{Introduction} In recent years, there has been a renaissance of interest in diffractive scattering. These diffractive processes are described by the Regge theory in terms of the Pomeron ($I\!\! P$) exchange\cite{pomeron}. The Pomeron carries quantum numbers of the vacuum, so it is a colorless entity in QCD language, which may lead to the ``rapidity gap" events in experiments. However, the nature of Pomeron and its reaction with hadrons remain a mystery. For a long time it had been understood that the dynamics of the ``soft pomeron'' is deeply tied to confinement. However, it has been realized now that how much can be learned about QCD from the wide variety of small-$x$ and hard diffractive processes, which are now under study experimentally. In Refs.\cite{th1,th2}, the diffractive $J/\psi$ and $\Upsilon$ production cross section have been formulated in photoproduction processes and in DIS processes in perturbative QCD. In the framework of perturbative QCD the Pomeron is represented by a pair of gluon in the color-singlet sate. This two-gluon exchange model can successfully describe the experimental results from HERA\cite{hera-ex}. On the other hand, as we know that there exist nonfactorization effects in the hard diffractive processes at hadron colliders \cite{preqcd,collins,soper,tev}. First, there is the so-called spectator effect\cite{soper}, which can change the probability of the diffractive hadron emerging from collisions intact. Practically, a suppression factor (or survive factor) ``$S_F$" is used to describe this effect\cite{survive}. Obviously, this suppression factor can not be calculated in perturbative QCD, which is now viewed as a nonperturbative parameter. Typically, the suppression factor $S_F$ is determined to be about $0.1$ at the energy scale of the Fermilab Tevatron\cite{tev}. Another nonfactorization effect discussed in literature is associated with the coherent diffractive processes at hadron colliders\cite{collins}, in which the whole Pomeron is induced in the hard scattering. It is proved in \cite{collins} that the existence of the leading twist coherent diffractive processes is associated with a breakdown of the QCD factorization theorem. Based on the success of the two-gluon exchange parametrization of the Pomeron model in the description of the diffractive photoproduction processes at $ep$ colliders\cite{th1,th2,hera-ex}, we may extend the applications of this model to calculate the diffractive processes at hadron colliders in perturbative QCD. Under this context, the Pomeron represented by a color-singlet two-gluon system emits from one hadron and interacts with another hadron in hard process, in which the two gluons are both involved (as shown in Fig.~1). The partonic process is plotted in Fig.2, where there are nine diagrams in the leading order of perturbative QCD. Therefore, these processes calculated in the two-gluon exchange model are just belong to the coherent diffractive processes in hadron collisions. Another important feature of the calculations of the diffractive processes in this model recently demonstrated is the sensitivity to the off-diagonal parton distribution function in the proton\cite{offd}. Using this two-gluon exchange model, we have calculated the diffractive $J/\psi$ production \cite{psi}, charm jet production\cite{charm}, massive muon pair and $W$ boson productions\cite{dy} in hadron collisions. These calculations show that we can explore much low $x$ (off-diagonal) gluon distribution function and study the coherent diffractive processes at hadron colliders through these processes. In this paper, we will calculate the light quark (massless) jet production at large transverse momentum in the coherent diffractive processes at hadron colliders by using the two-gluon exchange model. In the calculations of Refs.\cite{psi,charm,dy}, there always is a large mass scale associated with the production process. That is $M_\psi$ for $J/\psi$ production, $m_c$ for the charm jet production, $M^2$ for the massive muon production ($M^2$ is the invariant mass of the muon pair) and $M_W^2$ for $W$ boson production. However, in the light quark jet production process, there is no large mass scale. So, for the light quark jet production, the large transverse momentum is needed to guarantee the application of the perturbative QCD. Furthermore, the experience on the calculations of the diffractive di-quark photoproduction\cite{zaka} shows that the light quark jet production in the two-gluon exchange model has a distinctive feature that there is no contribution from the small $l_T^2$ region ($l_T^2<k_T^2$) in the integration of the amplitude over $l_T^2$. So, the expansion (in terms of $l_T^2/M_X^2$) method used in Refs.\cite{psi,charm,dy} can not be applied to the calculations of light quark jet production. In the following calculations, we will employ the helicity amplitude method to calculate the amplitude of the diffractive light quark jet production in hadron collisions. We will show that the production cross section is related to the differential (off-diagonal) gluon distribution function in the proton as that in the diffractive di-quark jet photoprodution process\cite{zaka}. (On the other hand, we note that the cross sections of the processes calculated in Refs.\cite{psi,charm,dy} are related to the integrated gluon distribution function in the proton). The diffractive production of heavy quark jet at hadron colliders has also been studied by using the two-gluon exchange model in Ref.\cite{levin}. However, their calculation method is very different from ours \footnote{For detailed discussions and comments, please see \cite{charm}}. In their calculations, they separated their diagrams into two parts, and called one part the coherent diffractive contribution to the heavy quark production. However, this separation can not guarantee the gauge invariance\cite{charm}. In our approach, we follow the definition of Ref.\cite{collins}, i.e., we call the process in which the whole Pomeron participants in the hard scattering process as the coherent diffractive process. Under this definition, all of the diagrams plotted in Fig.2 for the partonic process $gp\rightarrow q\bar qp$ contribute to the coherent diffractive production. The rest of the paper is organized as follows. In Sec.II, we will give the cross section formula for the partonic process $gp\rightarrow q\bar q p$ in the leading order of perturbative QCD, where we employ the helicity amplitude method to calculate the amplitude for this process. In Sec.III, we use this helicity amplitude method to recalculate the diffractive charm jet production process $gp\rightarrow c\bar c p$, by which we will reproduce the leading logarithmic approximation result for this process previously calculated in Ref.\cite{charm}. In Sec.IV, we will estimate the production rate of diffractive light quark jet at the Fermilab Tevatron by approximating the off-diagonal gluon distribution function by the usual diagonal gluon distribution function in the proton. And the conclusions will be given in Sec.V. \section{ The cross section formula for the partonic process} For the partonic process $gp\rightarrow q\bar qp$, in the leading order of perturbative QCD, there are nine diagrams shown in Fig.2. The two-gluon system coupled to the proton (antiproton) in Fig.2 is in a color-singlet state, which characterizes the diffractive processes in perturbative QCD. Due to the positive signature of these diagrams (color-singlet exchange), we know that the real part of the amplitude cancels out in the leading logarithmic approximation. To get the imaginary part of the amplitude, we must calculate the discontinuity represented by the crosses in each diagram of Fig.2. The first four diagrams of Fig.2 are the same as those calculated in the diffractive photoproduction processes. But, due to the existence of gluon-gluon interaction vertex in QCD, in the partonic process $gp\rightarrow q\bar q p$, there are additional five diagrams (Fig.2(5)-(9)). These five diagrams are needed for complete calculations in this order of QCD. In our calculations, we express the formulas in terms of the Sudakov variables. That is, every four-momenta $k_i$ are decomposed as, \begin{equation} k_i=\alpha_i q+\beta_i p+\vec{k}_{iT}, \end{equation} where $q$ and $p$ are the momenta of the incident gluon and the proton, $q^2=0$, $p^2=0$, and $2p\cdot q=W^2=s$. Here $s$ is the c.m. energy of the gluon-proton system, i.e., the invariant mass of the partonic process $gp\rightarrow q\bar q p$. $\alpha_i$ and $\beta_i$ are the momentum fractions of $q$ and $p$ respectively. $k_{iT}$ is the transverse momentum, which satisfies \begin{equation} k_{iT}\cdot q=0,~~~ k_{iT}\cdot p=0. \end{equation} All of the Sudakov variables for every momentum are determined by using the on-shell conditions of the momenta represented by the external lines and the crossed lines in the diagram. The calculations of these Sudakov variables are similar to those in the diffractive charm jet production process $gp\rightarrow c\bar c p$. And all of these Sudakov variables for the process $gp\rightarrow q\bar qp$ of the light quark jet production can take their values from the corresponding formulas in Ref.\cite{charm} after taking $m_c\rightarrow 0$. For convenience, we list all of the Sudakov variables for the diffractive process $gp\rightarrow q\bar qp$ in the following. For the momentum $u$, we have \begin{equation} \alpha_u=0,~~\beta_u=x_{I\! P}=\frac{M_X^2}{s},~~u_T^2=t=0, \end{equation} where $M_X^2$ is the invariant mass squared of the diffractive final state including the light quark and antiquark jets. For the high energy diffractive process, we know that $M_x^2\ll s$, so we have $\beta _u$ ($x_{I\! P}$) as a small parameter. For the momentum $k$, \begin{equation} \label{ak} \alpha_k(1+\alpha_k)=-\frac{k_T^2}{M_X^2},~~\beta_k=-\alpha_k\beta_u, \end{equation} where $k_T$ is the transverse momentum of the out going quark jet. For the loop mentum $l$, \begin{eqnarray} \nonumber \alpha_l&=&-\frac{l_T^2}{s},\\ \nonumber \beta_l&=&\frac{2(k_T,l_T)-l_T^2}{\alpha_ks},~~~{\rm for~Diag.}1,~3,~5,\\ \nonumber &=&\frac{2(k_T,l_T)+l_T^2}{(1+\alpha_k)s},~~~{\rm for~Diag.}2,~4,~6,\\ &=&-\frac{M_X^2-l_T^2}{s},~~~~~~~{\rm for~Diag.}7,~8,~9. \end{eqnarray} Using these Sudakov variables, we can give the cross section formula for the partonic process $gp\rightarrow q\bar q p$ as, \begin{equation} \label{xs} \frac{d\hat{\sigma}(gp\rightarrow q\bar qp)}{dt}|_{t=0}=\frac{dM_X^2d^2k_Td\alpha_k}{16\pi s^216\pi^3M_X^2} \delta(\alpha_k(1+\alpha_k)+\frac{k_T^2}{M_X^2})\sum \overline{|{\cal A}|}^2, \end{equation} where ${\cal A}$ is the amplitude of the process $gp\rightarrow q\bar qp$. We know that the real part of the amplitude ${\cal A}$ is zero, and the imaginary part of the amplitude ${\cal A}(gp\rightarrow q\bar qp)$ for each diagram of Fig.2 has the following general form, \begin{equation} \label{ima} {\rm Im}{\cal A}=C_F(T_{ij}^a)\int \frac{d^2l_T}{(l_T^2)^2}F\times\bar u _i(k+q)\Gamma_\mu v_j(u-k), \end{equation} where $C_F$ is the color factor for each diagram, and is the same as that for the $gp\rightarrow c\bar c p$ process\cite{charm}. $a$ is the color index of the incident gluon. $\Gamma_\mu$ represents some $\gamma$ matrices including one propagator. $F$ in the integral is defined as \begin{equation} F=\frac{3}{2s}g_s^3f(x',x^{\prime\prime};l_T^2), \end{equation} where \begin{equation} \label{offd1} f(x',x^{\prime\prime};l_T^2)=\frac{\partial G(x',x^{\prime\prime};l_T^2)}{\partial {\rm ln} l_T^2}, \end{equation} where the function $G(x',x^{\prime\prime};k_T^2)$ is the so-called off-diagonal gluon distribution function\cite{offd}. Here, $x'$ and $x^{\prime\prime}$ are the momentum fractions of the proton carried by the two gluons. It is expected that at small $x$, there is no big difference between the off-diagonal and the usual diagonal gluon densities\cite{off-diag}. So, in the following calculations, we estimate the production rate by approximating the off-diagonal gluon density by the usual diagonal gluon density, $G(x',x^{\prime\prime};Q^2)\approx xg(x,Q^2)$, where $x=x_{I\!\! P}=M_X^2/s$. In Ref.\cite{charm}, we calculate the amplitude (\ref{ima}) for the heavy quark diffractive production processes by expanding $\Gamma_\mu$ in terms of $l_T^2$. However, in the light quark jet production process $gp\rightarrow q\bar qp$, the expansion method is not yet valid. According to the result of \cite{charm}, the production cross section is proportional to the heavy quark mass. If we apply this formula to the light quark jet production, the cross section will be zero. That is to say, the expansion of the amplitude in terms of $l_T^2$, in which the large logarithmic contribution comes from the region of $l_T^2\ll M_X^2$, is not further suitable for the calculation of the cross section for massless light quark jet production. Furthermore, the experience of the calculation of the diffractive light quark jet photoproduction process $\gamma p\rightarrow q\bar qp$ (for $Q^2=0$) \cite{zaka} indicates that there is no contribution from the region of $l_T^2< k_T^2$ in the integration of the amplitude over $l_T^2$. In the hadroproduction process $gp\rightarrow q\bar qp$, the situation is the same. So, the expansion method, in which $l_T^2$ is taken as a small parameter, is not valid for the calculations of the massless quark production processes. In the following, we employ the helicity amplitude method\cite{ham} to calculate the amplitude Eq.(\ref{ima}). For the massless quark spinors, we define \begin{equation} u_\pm(p)=\frac{1}{\sqrt{2}}(1\pm\gamma_5)u(p). \end{equation} For the polarization vector of the incident gluon, which is transversely polarized, we choose, \begin{equation} \label{ev} e_\pm=\frac{1}{\sqrt{2}}(0,1,\pm i,0). \end{equation} The helicity amplitudes for the processes in which the polarized Dirac particles are involved have the following general forms\cite{ham}, \begin{equation} \label{ham} \bar u_\pm(p_f)Qv_\mp(p_i)=\frac{Tr[Q\not\! p_i\not\! n\not\! p_f(1\mp\gamma_5)]} {4\sqrt{(n\cdot p_i)(n\cdot p_f)}}, \end{equation} where $n$ is an arbitrary massless 4-vector, which is set to be $n=p$ in the following calculations. Using this formula (\ref{ham}), the calculations of the helicity amplitude ${\cal A}(\lambda(g),\lambda(q),\lambda({\bar q}))$ for the diffractive process $gp\rightarrow q\bar q p$ is straightforward. Here $\lambda$ are the corresponding helicities of the external gluon, quark and antiquark. In our calculations, we only take the leading order contribution, and neglect the higher order contributions which are proportional to $\beta_u=\frac{M_X^2}{s}$ because in the high energy diffractive processes we have $\beta_u\ll 1$. For the first four diagrams, to sum up together, the imaginary part of the amplitude ${\cal A}(\pm,+,-)$ is \begin{equation} \label{im1} {\rm Im}{\cal A}^{1234}(\pm,+,-)=\alpha_k^2(1+\alpha_k){\cal N}\times \int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2) (-\frac{2}{9}\frac{\vec{e}^{(\pm)}\cdot \vec{k}_T}{k_T^2}-\frac{1}{36}\frac{\vec{e}^{(\pm)}\cdot(\vec{k}_T-\vec{l}_T)}{(\vec{k}_T-\vec{l}_T)^2}), \end{equation} where $\frac{2}{9}$ and $-\frac{1}{36}$ are the color factors for Diag.1,4 and Diag.2,3 respectively, and ${\cal N}$ is defined as \begin{equation} \label{ne} {\cal N}=\frac{s}{\sqrt{-\alpha_k(1+\alpha_k)}}g_s^3T_{ij}^a. \end{equation} The other helicity amplitudes for the first four diagrams have the similar form as (\ref{im1}). The amplitude expression Eq.~(\ref{im1}) is the same as that for the photoproduction process $\gamma p\rightarrow q\bar q p$ previously calculated in Refs.\cite{th3,zaka} except the difference on the color factors. In the diffractive photoproduction process, the color factors of these four diagrams are the same (they are all $\frac{2}{9}$), while in hadroproduction process the color factors are no longer the same for these four diagrams. It is instructive to see what is the consequence of this difference. We know that the amplitude of the diffractive process in Eq.~(\ref{ima}) must be zero in the limit $l_T^2\rightarrow 0$. Otherwise, this will lead to a linear singularity when we perform the integration of the amplitude over $l_T^2$ due to existence of the factor $1/(l_T^2)^2$ in the integral of Eq.~(\ref{ima})\cite{charm}. This linear singularity is not proper in QCD calculations. So, we must first exam the amplitude behavior under the limit of $l_T^2\rightarrow 0$ for all the diffractive processes in the calculations using the two-gluon exchange model. From Eq.~(\ref{im1}), we can see that the amplitude for the diffractive photoproduction of di-quark jet process is exact zero at $l_T^2\rightarrow 0$. However, for the diffractive hadroproduction process $gp\rightarrow q\bar qp$ the amplitude for the first four diagrams is not exact zero in the limit $l_T^2\rightarrow 0$ due to the inequality of the color factors between them. So, for $gp\rightarrow q\bar qp$ process there must be other diagrams in this order of perturbative QCD calculation to cancel out the linear singularity which rises from the first four diagrams. The last five diagrams of Fig.2 are just for this purpose. For example, the contributions from Diags.5 and 8 are \begin{equation} {\rm Im}{\cal A}^{58}(\pm,+,-)=\alpha_k^2(1+\alpha_k){\cal N}\times \int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2) (-\frac{1+\alpha_k}{4}\frac{\vec{e}^{(\pm)}\cdot (\vec{k}_T-(1+\alpha_k)\vec{l}_T)}{(\vec{k}_T-(1+\alpha_k)^2\vec{l}_T)^2}), \end{equation} and the contributions from Diags.6 and 9 are \begin{equation} {\rm Im}{\cal A}^{69}(\pm,+,-)=\alpha_k^2(1+\alpha_k){\cal N}\times \int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2) (\frac{\alpha_k}{4}\frac{\vec{e}^{(\pm)}\cdot (\vec{k}_T-\alpha_k\vec{l}_T)}{(\vec{k}_T-\alpha_k^2\vec{l}_T)^2}), \end{equation} and the contribution from Diag.7 is \begin{equation} {\rm Im}{\cal A}^{7}(\pm,+,-)=\alpha_k^2(1+\alpha_k){\cal N}\times \int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2) (\frac{1}{2}\frac{\vec{e}^{(\pm)}\cdot \vec{k}_T}{k_T^2}). \end{equation} From the above results, we can see that the contributions from Diags.5-9 just cancel out the linear singularity which rises from the first four diagrams. Their total sum from the nine diagrams of Fig.2 is free of linear singularity now. Finally, by adding up all of the nine diagrams of Fig.2, the imaginary parts of the amplitudes for the following helicity sets are, \begin{eqnarray} \nonumber {\rm Im}{\cal A}(\pm,+,-)&=&\alpha_k^2(1+\alpha_k){\cal N}\times {\cal T}^{(\pm)},\\ {\rm Im}{\cal A}(\pm,-,+)&=&\alpha_k(1+\alpha_k)^2{\cal N}\times {\cal T}^{(\pm)}, \end{eqnarray} where \begin{eqnarray} \label{int2} \nonumber {\cal T}^{(\pm)}&=&\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)[(\frac{1}{2}-\frac{2}{9}) \frac{\vec{e}^{(\pm)}\cdot \vec{k}_T}{k_T^2}-\frac{1}{36}\frac{\vec{e}^{(\pm)}\cdot(\vec{k}_T-\vec{l}_T)}{(\vec{k}_T-\vec{l}_T)^2}\\ &&-\frac{1+\alpha_k}{4}\frac{\vec{e}^{(\pm)}\cdot(\vec{k}_T-(1+\alpha_k)\vec{l}_T)}{(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2} +\frac{\alpha_k}{4}\frac{\vec{e}^{(\pm)}\cdot(\vec{k}_T-\alpha_k\vec{l}_T)}{(\vec{k}_T-\alpha_k\vec{l}_T)^2}]. \end{eqnarray} And the amplitudes for the other helicity sets are zero in the light quark jet production process. From the above results, we can see that in the integration of the amplitude the linear singularity from different diagrams are canceled out by each other, which will guarantee there is no linear singularity in the total sum. Another feature of the above results for the amplitudes is the relation to the differential off-diagonal gluon distribution function $f(x',x'';l_T^2)$. To get the cross section for the diffractive process, we must perform the integration of Eq.~(\ref{int2}). However, as mentioned above that there is no big difference between the off-diagonal gluon distribution function and the usual gluon distribution at small $x$, so we can simplify the integration of (\ref{int2}) by approximating the differential off-diagonal gluon distribution function $f(x',x'';l_T^2)$ by the usual diagonal differential gluon distribution function $f_g(x;l_T^2)$. After integrating over the azimuth angle of $\vec{l}_T$, the integration ${\cal T}^{(\pm)}$ will then be \begin{equation} {\cal T}^{(\pm)}=\pi\frac{\vec{e}^{(\pm)}\cdot \vec{k}_T}{k_T^2}{\cal I}, \end{equation} where \begin{eqnarray} \label{inti} \nonumber {\cal I}&=&\int\frac{dl_T^2}{(l_T^2)^2}f_g(x;l_T^2)[\frac{1}{36}(\frac{1}{2}-\frac{k_T^2-l_T^2}{2|k_T^2-l_T^2|}) +\frac{1+\alpha_k}{4}(\frac{1}{2}-\frac{k_T^2-(1+\alpha_k)l_T^2}{2|k_T^2-(1+\alpha_k)l_T^2|})\\ &&-\frac{\alpha_k}{4}(\frac{1}{2}-\frac{k_T^2-\alpha_k l_T^2}{2|k_T^2-\alpha_k l_T^2|})]. \end{eqnarray} Comparing the above results with those of the photoproduction process $\gamma p\rightarrow q\bar q p$\cite{th3,zaka}, we find that the amplitude formula for the diffractive light quark jet hadroproduction process $gp\rightarrow q\bar qp$ is much more complicated. However, the basic structure of the amplitude, especially the expression for the integration ${\cal I}$ is similar to that for the photoproduction process. In the integration of (\ref{inti}), if $l_t^2<k_T^2$ the first term of the integration over $l_T^2$ will be zero; if $l_t^2<k_T^2/(1+\alpha_k)^2$ the second term will be zero; if $l_t^2<k_T^2/\alpha_k^2$ the third term will be zero. So, the dominant regions contributing to the three integration terms are $l_t^2\sim k_T^2$, $l_t^2\sim k_T^2/(1+\alpha_k)^2$, and $l_t^2\sim k_T^2/\alpha_k^2$ respectively. Approximately, by ignoring some evolution effects of the differential gluon distribution function $f_g(x;l_T^2)$ in the above dominant integration regions, we get the following results for the integration ${\cal I}$, \begin{equation} \label{i1} {\cal I}=\frac{1}{k_T^2}[\frac{1}{36}f_g(x;k_T^2)+\frac{(1+\alpha_k)^3}{4}f_g(x;\frac{k_T^2}{(1+\alpha_k)^2}) -\frac{\alpha_k^3}{4}f_g(x;\frac{k_T^2}{\alpha_k^2})]. \end{equation} Obtained the formula for the integration ${\cal I}$, the amplitude squared for the partonic process $gp\rightarrow q\bar qp$ will be reduced to, after averaging over the spin and color degrees of freedom, \begin{equation} \overline{|{\cal A}|}^2=\frac{9}{4}\alpha_s^3(4\pi)^3\pi^2s^2\frac{|{\cal I}|^2}{M_X^2} (1-\frac{2k_T^2}{M_X^2}). \end{equation} And the cross section for the partonic process $gp\rightarrow q\bar qp$ is \begin{equation} \label{xsp} \frac{d\hat\sigma(gp\rightarrow q\bar qp)}{dt}|_{t=0}=\int_{M_X^4>4k_T^2}dM_X^2dk_T^2 \frac{9\alpha_s^3\pi^2}{8(M_X^2)^2}\frac{1}{\sqrt{1-\frac{4k_T^2}{M_X^2}}} (1-\frac{2k_T^2}{M_X^2})|{\cal I}|^2. \end{equation} The integral bound $M_X^2>4k_T^2$ above shows that the dominant contribution of the integration over $M_X^2$ comes from the region of $M_X^2\sim 4k_T^2$. Using Eq.~(\ref{ak}), this indicates that in this dominant region $\alpha_k$ is of order of 1. So, in the integration ${\cal I}$ the differential gluon distribution function $f_g(x;Q^2)$ of the three terms can approximately take their values at the same scale of $Q^2=k_T^2$. That is, the integration ${\cal I}$ is then simplified to \begin{equation} \label{i2} {\cal I}\approx \frac{10M_X^2-27k_T^2}{36M_X^2}f_g(x;k_T^2). \end{equation} Numerical calculations show that there is little difference (within $10\%$ for $k_T>5~GeV$) between the cross sections by using these two different parametrizations of ${\cal I}$, Eq.~(\ref{i1}) and Eq.~(\ref{i2}). So, in Sec.IV, we use Eqs.~(\ref{xsp}) and (\ref{i2}) to estimate the diffractive light quark jet production rate at the Fermilab Tevatron. \section{Recalculate the heavy quark jet production using the helicity amplitude method} For a crossing check, in this section we will recalculate the diffractive heavy quark jet production at hadron colliders by using the helicity amplitude method. In Ref.\cite{charm}, we have calculated this process in the leading logarithmic approximation of QCD, where we expanded the amplitude in terms of $l_T^2$. Now, if we use the helicity amplitude method, we donot need to use the expansion method for the $\Gamma_\mu$ factor in Eq.~(\ref{ima}) as in \cite{charm}. We can firstly calculate the amplitude explicitly by using the helicity amplitude method. However, for the massive fermion, the amplitude formula is more complicated. Following Ref.\cite{ham1}, we first define the basic spinors $u_\pm(k_0)$ as, \begin{eqnarray} \label{k1} \nonumber u_+(k_0)=\not\! k_1u_-(k_0),\\ u_-(k_0)\bar u_-(k_0)=\frac{1}{2}(1-\gamma_5)\not\! k_0, \end{eqnarray} where the momenta $k_0$ and $k_1$ satisfy the followoing relations, \begin{equation} \label{k2} k_0\cdot k_0=0,~~~k_1\cdot k_1=-1,~~~k_0\cdot k_1=0. \end{equation} Using Eqs.~(\ref{k1}) and (\ref{k2}), we can easily find that the spinor $u_+(k_0)$ satisfies \begin{equation} u_+(k_0)\bar u_+(k_0)=\frac{1}{2}(1+\gamma_5)\not\! k_0. \end{equation} Provided the basic spinors, we then express any spinors $u(p_i)$ in terms of the basic ones,\cite{ham1} \begin{equation} u_\pm(p_i)=\frac{(\not\! p_i +m_i)u_\pm(k_0)}{\sqrt{2p_i\cdot k_0}}. \end{equation} It is easily checked that these spinors satisfy Dirac's equations. Now, for the massive fermions, the helicity amplitudes for the processes involving Dirac particles have the following general forms, \begin{eqnarray} \label{ham1} \nonumber \bar u_+(p_f)Qv_-(p_i)&=&\frac{Tr[Q(\not\! p_i-m_i)(1-\gamma_5)\not\! k_0(\not\! p_f+m_f)} {4\sqrt{(n\cdot p_i)(n\cdot p_f)}},\\ \nonumber \bar u_-(p_f)Qv_+(p_i)&=&\frac{Tr[Q(\not\! p_i-m_i)(1+\gamma_5)\not\! k_0(\not\! p_f+m_f)} {4\sqrt{(n\cdot p_i)(n\cdot p_f)}},\\ \nonumber \bar u_+(p_f)Qv_+(p_i)&=&\frac{Tr[Q(\not\! p_i-m_i)(1-\gamma_5)\not\! k_1\not\! k_0(\not\! p_f+m_f)} {4\sqrt{(n\cdot p_i)(n\cdot p_f)}},\\ \bar u_-(p_f)Qv_-(p_i)&=&\frac{Tr[Q(\not\! p_i-m_i)\not\! k_1(1-\gamma_5)\not\! k_0(\not\! p_f+m_f)} {4\sqrt{(n\cdot p_i)(n\cdot p_f)}}, \end{eqnarray} where $m_i$ and $m_f$ are the masses for the momenta $p_i$ and $p_f$ respectively, where $p_i^2=m_i^2,~p_f^2=m_f^2$. From the above equations, we can see that for the massless fermions ($m_i=m_f=0$) the formula Eq.~(\ref{ham1}) will then turn back to the formula Eq.~(\ref{ham}). To calculate the imaginary part of the amplitude Eq.~(\ref{ima}) for the partonic process $gp\to c\bar c p$, a convenient choice for the momenta $k_0$ and $k_1$ is, \begin{equation} k_0=p,~~~~k_1=e, \end{equation} where the vector $e$ is the polarization vector for the incident gluon defined in Eq.~(\ref{ev}). Using the formula Eq.~(\ref{ham1}) and the above choice for the momenta $k_0$ and $k_1$, the helicity amplitudes for Eq.~(\ref{ima}) will then be, \begin{eqnarray} \nonumber {\rm Im}{\cal A}(\pm,+,-)&=&\alpha_k^2(1+\alpha_k){\cal N}\times {\cal T}_c^{(\pm)},\\ {\rm Im}{\cal A}(\pm,-,+)&=&\alpha_k(1+\alpha_k)^2{\cal N}\times {\cal T}_c^{(\pm)},\\ {\rm Im}{\cal A}(\pm,+,+)&=&{\rm Im}{\cal A}(\pm,-,-)=\alpha_k(1+\alpha_k){\cal N}\times \frac{\pi m_c}{2}{\cal I}'_c, \end{eqnarray} where ${\cal N}$ is the same as in Eq.~(\ref{ne}), and the integrations ${\cal T}_c^{(\pm)}$ and ${\cal I}'_c$ are defined as \begin{eqnarray} \label{int3} \nonumber {\cal T}_c^{(\pm)}&=&\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)[(\frac{1}{2}-\frac{2}{9}) \frac{\vec{e}^{(\pm)}\cdot \vec{k}_T}{k_T^2+m_c^2}-\frac{1}{36}\frac{\vec{e}^{(\pm)}\cdot(\vec{k}_T-\vec{l}_T)}{m_c^2+(\vec{k}_T-\vec{l}_T)^2}\\ &&-\frac{1+\alpha_k}{4}\frac{\vec{e}^{(\pm)}\cdot(\vec{k}_T-(1+\alpha_k)\vec{l}_T)}{m_c^2+(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2} +\frac{\alpha_k}{4}\frac{\vec{e}^{(\pm)}\cdot(\vec{k}_T-\alpha_k\vec{l}_T)}{m_c^2+(\vec{k}_T-\alpha_k\vec{l}_T)^2}],\\ \nonumber {\cal I}'_c&=&\frac{1}{\pi}\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)[(\frac{1}{2}-\frac{2}{9}) \frac{1}{k_T^2+m_c^2}-\frac{1}{36}\frac{1}{m_c^2+(\vec{k}_T-\vec{l}_T)^2}\\ &&-\frac{1+\alpha_k}{4}\frac{1}{m_c^2+(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2} +\frac{\alpha_k}{4}\frac{1}{m_c^2+(\vec{k}_T-\alpha_k\vec{l}_T)^2}]. \end{eqnarray} If we approximate the differential off-diagonal gluon distribution function $f(x',x'';l_T^2)$ by the usual diagonal differential gluon distribution function $f_g(x;l_T^2)$, the above integrations will then be reduced to, after integrating over the azimuth angle of $\vec{l}_T$, \begin{equation} {\cal T}_c^{(\pm)}=\pi\vec{e}^{(\pm)}\cdot \vec{k}_T{\cal I}_c, \end{equation} where \begin{eqnarray} \label{int4} \nonumber {\cal I}_c&=&\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f_g(x;l_T^2)[\frac{5}{18} \frac{1}{m_T^2}-\frac{5}{36}\frac{1}{k_T^2}-\frac{1}{72}\frac{k_T^2-m_c^2-l_T^2}{m_1^2}\\ &&-\frac{1+\alpha_k}{8}\frac{k_T^2-m_c^2-(1+\alpha_k)^2l_T^2}{m_2^2} +\frac{\alpha_k}{8}\frac{k_T^2-m_c^2-\alpha_k^2l_T^2}{m_3^2}], \end{eqnarray} where \begin{eqnarray} \nonumber m_T^2&=&k_T^2+m_c^2,~~m_1^2=\sqrt{(m_t^2+l_T^2)^2-k_T^2l_T^2},~~m_2^2=\sqrt{(m_t^2+(1+\alpha_k)^2l_T^2)^2-(1+\alpha_k)^2k_T^2l_T^2},\\ ~~m_3^2&=&\sqrt{(m_t^2+\alpha_k^2l_T^2)^2-\alpha_k^2k_T^2l_T^2}, \end{eqnarray} and \begin{equation} \label{int5} {\cal I}'_c=\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f_g(x;l_T^2)[\frac{5}{18} \frac{1}{m_T^2}-\frac{1}{36}\frac{1}{m_1^2}- \frac{1+\alpha_k}{4}\frac{1}{m_2^2} +\frac{\alpha_k}{4}\frac{1}{m_3^2}]. \end{equation} So, the amplitude squared for the partonic process $gp\rightarrow c\bar cp$ will then be reduced to, after averaging over the spin and color degrees of freedom, \begin{equation} \label{cam} \overline{|{\cal A}|}^2=\frac{9}{4}\alpha_s^3(4\pi)^3\pi^2s^2\frac{m_T^2}{M_X^2} [(1-\frac{2k_T^2}{M_X^2})k_T^2|{\cal I}_c|^2+m_c^2|{\cal I}'_c|^2]. \end{equation} From the above results, we can see that the integrals of Eqs.~(\ref{int4}) and (\ref{int5}) are proportional to $1/l_T^2$ in the limit of $l_T^2\rightarrow 0$. That is to say that there exist large logarithmic contributions from the integration region of $1/R^2_N\ll l_T^2\ll m_T^2$ for the integration over $l_T^2$ as in Ref.\cite{charm}. So, we can expand the integrals of Eqs.~(\ref{int4}) and (\ref{int5}) in terms of $l_T^2$ to get the leading logarithmic contribution to the amplitude. In the limit of $l_T^2\rightarrow 0$, the parameters $m_1^2$, $m_2^2$ and $m_3^2$ scale as, \begin{eqnarray} \nonumber \frac{1}{m_1^2}&\approx & \frac{1}{m_T^2}[1-\frac{m_c^2-k_T^2}{m_T^2}\frac{l_T^2}{m_T^2}],\\ \nonumber \frac{1}{m_2^2}&\approx & \frac{1}{m_T^2}[1-\frac{m_c^2-k_T^2}{m_T^2}\frac{(1+\alpha_k)^2l_T^2}{m_T^2}],\\ \frac{1}{m_3^2}&\approx & \frac{1}{m_T^2}[1-\frac{m_c^2-k_T^2}{m_T^2}\frac{\alpha_k^2l_T^2}{m_T^2}]. \end{eqnarray} Under this approximation, the integrations Eqs.~(\ref{int4}) and (\ref{int5}) will then be related to the integrated gluon distribution function $xg(x;Q^2)$, \begin{eqnarray} {\cal I}_c\approx \frac{2m_c^2}{36(m_T^2)^3M_X^2}(10M_X^2-27m_T^2)xg(x;m_T^2),\\ {\cal I'}_c\approx \frac{m_c^2-k_t^2}{36(m_T^2)^3M_X^2}(10M_X^2-27m_T^2)xg(x;m_T^2). \end{eqnarray} Substituting the above results into Eq.~(\ref{cam}), we can then reproduce the leading logarithmic approximation result for the diffractive charm jet production process at hadron colliders which has been calculated in\cite{charm}. \section{Numerical results} Provided with the cross section formula (\ref{xsp}) for the partonic process $gp\rightarrow q\bar q p$, we can calculate the cross section of the diffractive light quark jet production at the hadron level. However, as mentioned above, there exists nonfactorization effect caused by the spectator interactions in the hard diffractive processes in hadron collisions. Here, we use a suppression factor ${\cal F}_S$ to describe this nonfactorization effect in the hard diffractive processes at hadron colliders\cite{soper}. At the Tevatron, the value of ${\cal F}_S$ may be as small as ${\cal F}_S\approx 0.1$\cite{soper,tev}. That is to say, the total cross section of the diffractive processes at the Tevatron may be reduced down by an order of magnitude due to this nonfactorization effect. In the following numerical calculations, we adopt this suppression factor value to evaluate the diffractive production rate of light quark jet at the Fermilab Tevatron. The numerical results of the diffractive light quark jet production at the Fermilab Tevatron are plotted in Fig.3 and Fig.4. In our calculations, the scales for the parton distribution functions and the running coupling constant are both set to be $Q^2=k_T^2$. For the parton distribution functions, we choose the GRV NLO set \cite{grv}. In Fig.3, we plot the differential cross section $d\sigma/dt|_{t=0}$ as a function of the lower bound of the transverse momentum of the light quark jet, $k_{T{\rm min}}$. This figure shows that the cross section is sensitive to the transverse momentum cut $k_{T{\rm min}}$. The differential cross section decreases over four orders of magnitude as $k_{T{\rm min}}$ increases from $5~GeV$ to $15~GeV$. It is interesting to compare the cross section of the diffractive light quark jet production with that of the diffractive charm quark jet production\cite{charm}, which is also shown in Fig.3. The two curves in this figure show that the production rates of the light quark jet and the charm quark jet in the diffractive processes are in the same order of magnitude. However, we know that the cross section of diffractive heavy quark jet production is related to the integrated gluon distribution function, while the cross section of the light quark jet production is related to the differential gluon distribution function. In Fig.4, we plot the differential cross section $d\sigma/dt|_{t=0}$ as a function of the lower bound of the momentum fraction of the proton carried by the incident gluon $x_{1{\rm min}}$, where we set $k_{T{\rm min}}=5~GeV$. This figure shows that the dominant contribution comes from the region of $x_1\sim 10^{-2}-10^{-1}$. This property is the same as that of the diffractive charm jet production at the Tevatron\cite{charm}. \section{Conclusions} In this paper, we have calculated the diffractive light quark jet production at hadron colliders in perturbative QCD by using the two-gluon exchange model. We find that the production cross section is related to the squared of the differential gluon distribution function $\partial G(x;Q^2)/\partial ln Q^2$ at the scale of $Q^2\sim k_T^2$, where $k_T$ is the transverse momentum of the final state quark jet. For a crossing check, we also used the helicity amplitude method to calculate the diffractive charm jet production process at hadron colliders, by which we could reproduce the leading logarithmic approximation result for this process previously calculated in Ref.\cite{charm}. We have also compared the production rate of the light quark jet in the diffractive processes with that of the heavy quark jet production, and found that the production rates of these two processes are in the same order of magnitude. As we know, the large transverse momentum dijet production in the diffractive processes at hadron colliders is important to study the diffractive mechanism and the nature of the Pomeron. The CDF collaboration at the Fermilab Tevatron have reported some results on this process\cite{tev}. In this paper, under the two-gluon exchange model in perturbative QCD we have calculated the light quark jet production in the diffractive processes. In a forthcoming paper, we will calculate the diffractive gluon jet production in hadron collisions by using the two-gluon exchange model, which will complete the calculations of the diffractive dijet production in perturbative QCD. These results then can be used to compare with the experimental measurements to test the valid of the perturbative QCD description of the diffractive processes at hadron colliders and may be used to study the diffractive mechanism and the factorization broken effects. \acknowledgments This work was supported in part by the National Natural Science Foundation of China, the State Education Commission of China, and the State Commission of Science and Technology of China.
\section{Introduction} The identification of mesons beyond the conventional valence quark-antiquark $(Q\bar Q)$ pair configuration is one of the most challenging tasks in low-energy meson spectroscopy. \\ For instance, as a consequence of gluon self-couplings in QCD the usual $Q\bar Q$ spectrum is expected to be supplemented by mesonic states with pure gluonic composition, that is glueballs \cite{clos}.\\ Lattice QCD, in the quenched approximation, predicts the lightest glueball state $(G_0)$ to be a scalar $(J^{PC}=0^{++})$ lying in the mass range of 1.4 - 1.8 GeV \cite{bal,tep}. Since the observed isovector and isodoublet states $a_0(1450)$ and $K_0(1430)$ set the natural mass scale of the ground state scalar meson nonet, the isoscalar $Q\bar Q$ partners $n \bar n \equiv \frac{1}{\sqrt{2}}(u\bar{u}+d\bar{d})$ and $s \bar s$ are expected to lie in a mass region close to the predicted ground state glueball. The vicinity of masses for the pure $G_0$, $n\bar n$ and $s\bar s$ mesonic states suggests that the intrusion of the scalar glueball into the scalar $Q\bar Q$ meson spectrum is accompanied by significant mixing of $G_0$ with the isoscalar members of the $^3P_0$ nonet \cite{ams,weina}.\\ Several glueball-quarkonia mixing schemes \cite{ams,weina,anis,naris} have been proposed to account for this scenario and to reveal the $G_0/Q\bar Q$ nature of the scalar-isoscalar states $f_0(1370)$, $f_0(1500)$, established by Crystal Barrel at LEAR \cite{amsler} in proton-antiproton annihilation reactions, and the $f_0(1710)$ \cite{bugg1710}, which is the scalar component of the $f_{j=0,2}(1710)$ \cite{pdg}. In a simplified picture considered in Refs. \cite{ams,weina} the scalar glueball ground state $G_0$ mixes with the lowest lying $^3P_0$ isoscalar quarkonium states. Further admixture of the first radially excited $^3P_0$ nonet is introduced in Refs. \cite{anis,naris}. However, the restriction to the glueball and near lying $^3P_0$ quarkonia ground states is a good approximation provided that the excited glueball and quarkonia states are high enough in mass.\\ Quantitative predictions in the three-state mixing schemes of Refs. \cite{ams,weina} for the glueball-quarkonia content of $f_0(1370)$, $f_0(1500)$ and $f_0(1710)$ differ substantially due to orthogonal theoretical assumptions concerning the level ordering of the bare states before mixing. While Ref. \cite{ams} uses for the input masses of the bare states $M(n\bar n)< M(G_0)<M(s\bar s)$, in Ref. \cite{weina} the pure glueball lies above the scalar $Q\bar Q$ state. The difference of level ordering in the input bare masses leads to a substantially different $G_0/Q\bar Q$ content of the physical $f_0$ states, especially for $f_0(1500)$ and $f_0(1710)$. While in Ref. \cite{ams}, the glueball is distributed with nearly equal strength over the three $f_0$ states and $f_0(1710)$ has a large $s \bar s$ component, the analysis of Ref. \cite{weina} leads to a dominant $G_0$ structure in the $f_0(1710)$, while $f_0(1500)$ is a nearly pure $s\bar s$ state. The results for the $f_0(1370)$ are comparable in both approaches, with a large $n \bar n$ component residing in this state. The observed total width of the $f_0(1370)$ \cite{amsler} and its strong coupling to the $\pi\pi$ decay channel is in good agreement with naive quark model expectations \cite{kok,god} for a $n\bar n$ $^3P_0$ state and seems to confirm its dominant $n\bar n$ structure. Therefore, the main difference between the two mixing schemes is the amount of glueball and $s\bar s$ strength residing in the $f_0(1500)$ and $f_0(1710)$ states, respectively. A further test of the proposed mixing schemes, where the glueball ground state intrudes in the scalar quarkonium sector, is the analysis of the decay modes of the observed $f_0$ states. The decay of $f_0(1500)$ into two pseudoscalar mesons was studied in leading order of strong coupling QCD \cite{ams}. Neglecting possible gluonic components of the isoscalar mesons in the final state, the $ \pi\pi, K\bar K, \eta\eta$ and $ \eta\eta^{\prime} $decay modes are entirely driven by the quarkonia components of the $f_0(1500)$ state. In this approximation, results for the partial decay widths deduced from $SU(3)$ flavor symmetric couplings are quite similar for both mixing schemes and cannot be discriminated \cite{ams}, e.g. the observed weak $K\bar K$ decay mode is obtained in both schemes by the destructive interference between the $n\bar n$ and $s\bar s$ components. \\ In order to reveal the difference of the proposed three state mixing schemes, one should take into account the direct coupling of the $G_0$ component to the final state mesons. This mechanism occurs in the leading of strong coupling QCD \cite{ams} by the transition $G_0 \rightarrow G_0 G_0$, which can produce final states with isoscalar mesons $(\eta\eta, \eta\eta^\prime$ and $\sigma\sigma$, where $\sigma$ is the broad low lying $\pi\pi$ S-wave resonance) with possibly a sizable gluonic component. Independent of a possible small glueball component in the final state isoscalars is the decay process $G_0 \rightarrow (Q\bar Q) (Q\bar Q)$. This next-to-leading order transition can contribute to all two-body decay modes of the mixed $f_0$ states. An investigation \cite{mitja} of the two-body decay modes $\sigma\sigma, \rho\rho$ and $\pi^\ast(1300)\pi$ of the $f_0(1500)$, all leading to the $4\pi$ decay channel, indicates the possible presence of a sizable direct coupling of the gluonic component of the $f_0(1500)$ to the $4\pi$ decay channels. Neglecting a possible $G_0$ component in $\sigma$, the leading order decay mechanism predicts \cite{mitja} the hierarchy of branching ratios BR with $BR(\rho\rho)^{>}_{\sim} BR(\pi\pi) ^{>}_{\sim} BR(\sigma\sigma) > BR(\pi^*(1300)\pi)$, independent of any particular three-state mixing scheme. The results are in strong conflict with current Crystal Barrel data, where the $\rho\rho$ decay mode of the $f_0(1500)$ is weak or even absent \cite{thoma} and a stronger coupling to the $4\pi$ modes than to the $2\pi$ decay channel \cite{amsler} is required. \\ In the present work, going beyond the lowest order decay mechanism, we couple the $Q\bar Q$ components of the mixed $f_0$ states to the quarkonia components of the final state mesons. In the decay analysis we also take into account the transition $G_0 \rightarrow (Q\bar Q)~(Q\bar Q)$ occurring as a next-to-leading order decay mechanism deduced from strong coupling QCD \cite{ams}. Inclusion of the direct coupling of the $G_0$ component of the mixed $f_0$ states to the quarkonia content of the final state mesons will lead to contradictory physical consequences in the proposed three-state mixing schemes of Refs. \cite{ams,weina}, in particular for the physical state $f_0(1710)$. Because of the shortcomings of these proposed mixing models, we undertake to extract a three-state mixing scheme based on a fit to the detailed experimental data of $f_0(1500)$ two-body decays and give the predictions for partial decay widths of the partner states $f_0(1370)$ and $f_0(1710)$.\\ This paper is organised as follows. In Sec. \ref{mixing} we introduce the three-state mixing scheme of the lowest lying scalar glueball with the quarkonia states and the decay dynamics as suggested by the strong coupling limit of QCD. In Sec. \ref{quark} we develop the formalism for the decay of the $f_0$ states by its $Q \bar Q$ components in the framework of the pair creation or $^3P_0$ model. In Sec. \ref{gluon} we construct a digluonium wave function based on cavity QCD for the effective description of the scalar glueball state. Furthermore, we indicate the transition mechanism for the decay of the $G_0$ component into $(Q\bar Q)~(Q\bar Q)$. The results are presented in Sec. \ref{clo-wein}; we first show how distinct decay patterns result from the unmixed $Q\bar Q$ and $G_0$ components of the $f_0$ states. We then give partial widths for the two-body decay modes of the $f_0$ states in the mixing schemes of Refs. \cite{ams,weina} and confront the results with experiment. Finally, we present a three-state mixing scheme deduced from a detailed fit to the experimental data for $f_0(1500)$ decays and give predictions for the partial decay widths of $f_0(1370)$ and $f_0(1710)$, providing thus a sensitive test of the proposed mixing scheme. A summary and conclusions are given in Sec. \ref{summary}. \newpage \section{Three-state mixing and decay of scalar-isoscalar states in strong coupling QCD} \label{mixing} In the lattice formulation of the strong coupling limit $(g \rightarrow \infty)$ of QCD \cite{ams,kok}, glueballs and $Q\bar Q$ mesons are noninteracting eigenstates of the QCD Hamiltonian. Departure from this limit causes mixing and decay of the strong coupling eigenstates \cite{ams,kok}.\\ To lowest order, the mixing of the scalar glueball $G_0$ and the $^3P_0$ quarkonia states $n\bar n=1/\sqrt(2)~(u\bar u+d\bar d)$ and $s\bar s$ can be introduced by the interaction Hamiltonian $H_I$ \cite{weina} \begin{equation} (~H_I~)= \left( \begin{array}{ccc} m_{G_0} & z & \sqrt{2} z\\ z & m_{s\bar s}& 0\\ \sqrt{2}z & 0 & m_{n\bar n} \end{array} \right), \end{equation} where $z=<G_0|H_I|s\bar s>=<G_0|H_I|n\bar n>/\sqrt{2}$ represents the flavor independent mixing strength between the glueball and quarkonium states. The masses of the bare states $i=G_0, s\bar s, n\bar n$ before mixing are denoted by $m_i$. Deviation from the $SU(3)$ flavor symmetry is known to be small \cite{weinb} and quarkonia mixing is assumed to be a higher order perturbation in the strong coupling eigenstates $|G_0>,~|s\bar s>$ and $|n\bar n>$ \cite{ams}. All parameters of $H_I$ can be taken to be real and positive.\\ $H_I$ possesses three eigenstates $|f_0(M)>$ with physical masses M , which are are given by the linear combinations: \begin{equation} \label{mix} |f_0(M)> = a_{n\bar n}|n\bar n>+a_{s\bar s}|s\bar s>+a_{G_0}|G_0>, \end{equation} with $a_{n\bar n}^2+a_{s\bar s}^2+a_{G_0}^2=1$. Choosing the bare mass $m_{n\bar n}$ to be equal to that of the observed isovector state $a_0(1450)$ \cite{pdg}, the remaining parameters in $H_I$ can be adjusted to give the masses of the observed resonances $f_0(1370)$, $f_0(1500)$ and $f_0(1710)$, which fixes the quarkonia-glueball composition of the physical states \cite{weina}. In Ref. \cite{ams} it is assumed that the level spacings of the bare states fulfil the relation $(m_{G_0}-m_{n\bar n})/2=m_{s\bar s}-m_{G_0}=z$, which again determines the state vector of the physical $f_0(M)$ states. In this case $f_0(1500)$ is degenerate in mass with the $G_0$ state and the physical masses M of the mixing partners $f_0(1370)$ and $f_0(1710)$ depend on the mixing strength z. \\ However, there are sizable uncertainties in these procedures, particularly since the mass of the $f_0(1370)$ is poorly determined, ranging from 1200 - 1500 MeV \cite{pdg} and the exact level spacing of the bare states in unknown. For these reasons we will also present an alternative strategy by assuming that $f_0(1500)$ is one of the eigenstates of $H_I$ and by extracting the mixing coefficients $a_i$ $(i= n\bar n, s\bar s, G_0)$ directly from the experimental two-body decay data. This will allow us to give the bare masses $m_i$ entering in $H_I$ of Eq. (1), in terms of the mixing strength z, which in turn is restricted by the physical masses of the partners of the $f_0(1500)$. The resulting glueball-quarkonia composition of these partner states , however, is completely independent of z, solely determined by the mixing coefficients of the $f_0(1500)$. \\ In the minimal constituent picture $G_0$ consists of two gluons bound in a color singlet configuration with $J^{PC}=0^{++}$. For finite $g$ the mixed $|f_0(M)>$ decays in lowest order by its $G_0$ and $Q\bar Q$ components. The relevant diagrams are shown in Figs. \ref{glueb} and \ref{quarkonia}. The transition $G_0 \rightarrow G_0~ G_0$ of Fig. 1 couples the $G_0$ component of the physical state $|f_0(M)>$ to the $G_0$ component of the final state isoscalar mesons. The constituent gluons of the initial state decay into two pairs of low energy gluons which involves two three-gluon vertices in the decay mechanism. We will neglect this transition in our decay analysis by assuming that the considered final state isoscalar mesons do not contain a $G_0$ component. In principle, the transition of Fig. 1 can contribute to the decay modes $f_0(M) \rightarrow \eta\eta, \eta\eta^\prime$ and $\sigma\sigma$, depending on the unknown overlap of the $\eta, \eta^\prime$ and $\sigma$ with the corresponding glueball states. Current evidence for the relevance of such a process involving the pseudoscalars $\eta$ and $\eta^\prime$ is not compelling, as for example discussed in Ref. \cite{ams}. For the scalar $\sigma$ meson a large coupling to the glueball state is not excluded \cite{ams,pdg}. Therefore, we should keep in mind that the $G_0 \rightarrow G_0~G_0$ transition could induce some changes for the $\eta\eta, \eta\eta^\prime$ and $\sigma\sigma$ decay modes of the $f_0$ states.\\ The leading order decay mechanism $Q\bar Q \rightarrow (Q\bar Q)~(Q\bar Q)$, illustrated in Fig. \ref{quarkonia}, leads to the decay of $|f_0(M)>$ by its quarkonia components $Q \bar Q$ and is familiar from the OZI-allowed meson decay \cite{kok,mitja,meson}. This process will be described in Sec. \ref{quark} in the framework of the $^3P_0$ pair creation model \cite{le}.\\ Here we also take into account the next to leading order decay mechanism $G_0 \rightarrow (Q\bar Q)~(Q\bar Q)$ as indicated in Fig. \ref{gluea}. In this mechanism the $G_0$ component of the physical $f_0(M)$ states couples to the final state $Q\bar Q$ mesons by conversion of the constituent gluons into quark-antiquark pairs. Sec. \ref{gluon} is devoted to the description of this process.\\ Neglecting a possible $G_0$ component in the final state isoscalar mesons, the partial decay width $\Gamma_{f_0 \rightarrow BC}$ for the decay of $|f_0(M)>$ into a two-meson state B C can be written as \begin{eqnarray} \Gamma_{f_0 \rightarrow BC}& = & 2~\pi~ K~ \frac{E_BE_C}{M_{f_0}}~ \sum_{l_{BC}}~ \int d\Omega_K~ |T_{f_0 \rightarrow BC}^{(l_{BC})}|^2 \\ & = & \sum_{l_{BC}}~|\lambda~M_{Q\bar Q \rightarrow BC}^{(l_{BC})}~ +~\gamma^2~M_{G_0 \rightarrow BC}^{(l_{BC})}|^2 ~, \nonumber \end{eqnarray} where $M_{Q\bar Q \rightarrow BC}^{(l_{BC})}$ and $M_{G_0 \rightarrow BC}^{(l_{BC})}$ denote the transition amplitudes including phase space of the respective decay mechanism. The sum extends over the relative orbital angular momentum $l_{BC}$ between the mesons B and C. With the decay momentum $|\vec K|=K$, $E_i=\sqrt{M_i^2+\vec K^2}$ is the energy of the final state mesons $i=B, C$ with mass $M_i$. The strengths of the quark-gluon coupling and the pair creation amplitude in the $^3P_0$ model are factorised out and denoted by $\gamma$ and $\lambda$, respectively. The relative phase between the two transition amplitudes is fixed by choosing $M_{Q\bar Q \rightarrow BC}^{(l_{BC})}$ and $M_{G_0 \rightarrow BC}^{(l_{BC})}$ to be real and the ratio $\gamma^2/\lambda$ to be complex with \begin{equation} \frac{\gamma^2}{\lambda}~\equiv~\kappa~e^{i\phi}. \end{equation} In this case $\Gamma_{f_0 \rightarrow BC}$ is given by \begin{eqnarray} \Gamma_{f_0 \rightarrow BC} & = & \lambda^2 \left\{ M^2_{Q\bar Q \rightarrow BC} + \kappa^2~ M^2_{G_0 \rightarrow BC} + 2~\kappa ~cos\phi~ \sum_{l_{BC}}~ M^{(l_{BC})}_{Q\bar Q \rightarrow BC} M^{(l_{BC})}_{G_0 \rightarrow BC} \right\}. \end{eqnarray} with the summation over $l_{BC}$ included in the definition of $M^2_{Q\bar Q (G_0) \rightarrow BC}$. Finally, in order to take into account properly the available phase space for the decay $f_0 \rightarrow BC$ we average over the mass spectrum $f(m)$ of broad mesons: \begin{eqnarray} \Gamma_{f_0 \rightarrow BC} & = & \int dm_{f_0} dm_B dm_C ~ \Gamma_{f_0 \rightarrow BC}(m_{f_0}, m_B, m_C) f(m_{f_0}) f(m_B) f(m_C) \\ f(m) & \propto & \frac{(\Gamma_i/2)^2}{(m-M_i)^2+(\Gamma_i/2)^2} \nonumber \end{eqnarray} with a threshold cutoff as in Ref. \cite{mary}. The individual masses $M_i$ and widths $\Gamma_i$ of the resonances are taken from the Particle Data Group.\\ The mass distribution of the scalar-isoscalar $\sigma$ meson is parameterised with $M_\sigma =760~MeV$ and $\Gamma_\sigma=640~MeV$ \cite{nag}. For the $\pi^*(1300)$ we use the resonance width of $\Gamma_{\pi^*(1300)}=200~MeV$ \cite{ulrike}. The relationship between $f_0(1370)$ and the broad structure around $1100~MeV$ \cite{amsler} called $f_0(400-1200)$ \cite{pdg} is not entirely clear. If the two states are manifestations of a single object, then $\Gamma_{f_0(1370)}\approx 700~MeV$ \cite{amsler}. For two independent states one has $\Gamma_{f_0(1370)}=351\pm41~MeV, M_{f_0(1370)}=1360\pm23~MeV$ \cite{amsler} and $\Gamma_{f_0(1370)}=230\pm15~MeV, M_{f_0(1370)}=1300\pm15~MeV$ \cite{bugga}. \newpage \section{Decay of the quarkonia components} \label{quark} The decay of the quarkonia components of the $|f_0(M)>$ states into the final state mesons B and C is calculated in the non-relativistic $^3P_0$ or pair creation model \cite{le}. This model has a solid foundation in strong coupling QCD \cite{kok,pat,dosch}. It describes the OZI-allowed decay $Q\bar Q \rightarrow BC$ by the creation of a $Q\bar Q$ pair with quantum numbers $(I^G(J^{PC})=0^+(0^{++}))$ out of the hadronic vacuum, as indicated in Fig. \ref{quarkonia}. In the evaluation one also has to consider a second diagram where the quark 3 (4) goes into the meson C (B). The constituent quarks of the initial state are spectators in the transition. The strength of this transition is governed by the dimensionless constant $\lambda$, which is related to the pair creation probability. From detailed fits to tensor meson decay \cite{ams,thomas} it is concluded that the pair creation mechanism is flavor independent.\\ The nonperturbative $Q\bar Q$ $^3P_0$ vertex can be defined as \begin{equation} V_{^3P_0}^{(43)}= \lambda~ \delta ( {\vec{p}} _{4}+ {\vec{p}} _{3}) \left[ {\cal Y}_{1\mu}^\ast( {\vec{p}} _{4} - {\vec{p}} _{3}) \otimes \sigma_{-\mu}^{(43)\dagger} \right]_{00} {\mathbf{1}} _F^{(43)} {\mathbf{1}} _C^{(43)}~, \end{equation} where $\vec{p}_{4(3)}$ are the momenta of the quark (antiquark) 4 (3) and ${\cal Y}_1(\vec p)=|\vec p|~Y_1(\hat p)$. The identity operators $\mathbf 1_F$ and $\mathbf 1_C$ project onto singlet states in flavor (F) and color (C) space of the created $Q\bar Q$ pair (43).\\ Using the harmonic oscillator ansatz $\Psi_{n=0,l=1}(\vec{p_1},\vec{p_2})$, the wave function of the $Q\bar Q$ component of the mixed $f_0$ can be expressed in its centre-of-momentum frame as: \begin{eqnarray} \label{a} \Psi_{f_0}(\vec{p}_1,\vec{p}_2) & = & \delta ( {\vec{p}}_1 + {\vec{p}}_2 )~ \left[\chi_{S=1}(12)\otimes \Psi_{n=0,l=1} ({\vec{p}}_1,{\vec{p}}_2)\right]_{J=0}~ \chi_F^{f_0}(12)~\chi_C^{(12)} \nonumber\\ & = & \left[ \frac{2R_{Q\bar Q}^5}{3\sqrt{\pi}} \right]^{\frac{1}{2}} \delta ( {\vec{p}}_1 + {\vec{p}}_2 ) \exp \left\{ - \frac{1}{8} R_{Q\bar Q}^2 ( {\vec{p}}_1-{\vec{p}}_2)^2 \right\}\\ & & \cdot \left[ \chi_{S=1}(12) \otimes {\cal Y}_{l=1}({\vec{p}}_1-{\vec{p}}_2) \right]_{J=0} \chi_{F}^{f_0}(12) \chi_C^{(12)} ~, \nonumber \end{eqnarray} where $R_{Q\bar Q}$ is the size parameter for the $Q\bar Q$ component. The intrinsic spin $\chi_{S=1}(12)$ of the $Q\bar Q$ pair (12) is coupled with relative angular momentum $l=1$ to the $J^{PC}=0^{++}$ scalar ground state ($n=0$) and $\chi_C^{(12)}$ is the color singlet wave function. The $SU(3)$ flavor part $\chi^{f_0}_F(12)$ is given by \begin{equation} \chi^{f_0}_F(12)~=~a_{n\bar n}~|n_1\bar n_2>~+~a_{s\bar s}~|s_1\bar s_2>, \end{equation} with the quarkonia mixing coefficients $a_i$ ($i=n\bar n, s\bar s$) of the three-state mixing scheme introduced in Sec. \ref{mixing}. For the final state mesons $i~ (i=B,C)$ with momentum $\vec P_i$, total spin $J_i$, relative orbital angular momentum $l_i$ and intrinsic spin $S_i$ we use the wave functions \begin{equation} \Psi_{i}(\vec p_l,\vec p_m)~=\delta ( {\vec{p}}_l + {\vec{p}}_m-\vec P_i )~ [\chi_{S_i}(lm)\otimes\Psi_{n_il_i} (\vec{p_l},\vec{p_m})]_{J_i}~\chi^i_F(lm)~\chi_C^{(lm)}, \end{equation} where $\vec p_{l(m)}$ are internal quark momenta for the quarks $l(m)$. Explicit forms for the harmonic oscillator wave functions $\Psi_{n_il_i}$ used can be found in Appendix A. \\ The transition amplitude $M_{Q\bar Q \rightarrow BC}^{(l_{BC})}$, as defined by Eq. (3), separates into a flavor part $M^F_{Q\bar Q \rightarrow BC}$ \begin{equation} M^F_{Q\bar Q \rightarrow BC}~=~<\chi^B_F(13)\chi^C_F(42)|{\mathbf{1}}_F^{(43)}|\chi_F^{f_0}(12)> \end{equation} and the spin-spatial-color part $M^{(l_{BC})SSC}_{Q\bar Q \rightarrow BC}$ \begin{eqnarray} M_{Q\bar Q \rightarrow BC}^{(l_{BC})SSC}& = & \sqrt{2\pi K}~ \sqrt{\frac{E_BE_C}{M_{f_0}}}~ \frac{1}{\sqrt{3}}\int\prod_{i=1,...4}d\vec {p_i} ~\delta(\vec{p_4}+\vec{p_3})~ \delta(\vec p_1+\vec p_2) \\ & & \delta(\vec p_1+\vec p_3-\vec K)~ \delta(\vec p_2+\vec p_4+\vec K)~ [\chi_{S_B}^{(13)}\otimes \Psi_{n_Bl_B}(\vec{p_1},\vec{p_3})]^\dagger \nonumber\\ & & [\chi_{S_C}^{(42)}\otimes \Psi_{n_Cl_C}(\vec{p_4},\vec{p_2})]^\dagger \left[ {\cal Y}_{1\mu}^\ast( {\vec{p}} _{4} - {\vec{p}} _{3}) \otimes \sigma_{-\mu}^{(43)\dagger} \right]_{00} [\chi_{S=1}^{(12)}\otimes \Psi_{n=0l=1}(\vec{p_1},\vec{p_2})]_{J=0} \nonumber \end{eqnarray} with the decay momentum $\vec K$ and the color matrix element $<\chi_C^{(13)}\chi_C^{(42)}|{\mathbf{1}}_C|\chi_C^{(12)}>~=~\sqrt{1/3}$. \\ The amplitude $M_{Q\bar Q \rightarrow BC}^{(l_{BC})SSC}$ is evaluated analytically. Since there exists exhaustive literature \cite{kok,meson,le} on the $^3P_0$ model we just cite the results in Appendix B. Values for the various flavor matrix elements are given in Table \ref{tab1}. The results of Table \ref{tab1} were obtained by assuming $\chi^{\omega}_F=~|n\bar n>$. Thereby we have used the following definitions for the flavor content of the $\eta$ and $\eta^\prime$ mesons: \begin{eqnarray} \psi_F^{\eta}& = & \alpha_{PS}~|n\bar n>~-~\beta_{PS}~|s\bar s> \\ \psi_F^{\eta^\prime}& = & \beta_{PS}~|n\bar s>~+~\alpha_{PS}~|s\bar s> \nonumber . \end{eqnarray} The quantities $\alpha_{PS}$ and $\beta_{PS}$ are simply related to the pseudoscalar mixing angle $\theta_{P}$ as defined by the Particle Data Group \cite{pdg}. For numerical results we use the value of $\theta_{P}=-~17.3^o$ \cite{ps} or equivalently $\alpha_{PS}~=~0.794$, $\beta_{PS}~=~0.608$. \\ The partial decay widths for the quarkonia decay $Q\bar Q \rightarrow BC$ are defined as in Eq. (5) and depend on the pair creation amplitude $\lambda$ and the quarkonia size parameters $R_{Q\bar Q}$ and $R_B=R_C$ for the decay channels considered here.\\ Exhaustive experimental data exist for the hadronic decays of the ground state tensor meson nonet and the strategy is to extract the meson size parameters combined with the $^3P_0$ strength from a full scale decay analysis. A recent fit to tensor meson decays \cite{thomas} yields $\lambda^2=2.11$ and $R_B=3.65~GeV^{-1}$ for nonstrange S-wave mesons and strange tensor mesons, suggesting $R_{a_1}=R_{Q\bar Q}=3.65~GeV^{-1}$. S-wave mesons containing strange quarks have an effectively reduced size with $R_B=2.48~GeV^{-1}$ for $\eta$ and $\eta^\prime$ and $R_K=2.82~GeV^{-1}$ for K. No flavor symmetry breaking at the $^3P_0$ vertex was obtained, hence $\lambda$ is independent of the created $Q\bar Q$ flavor. Due to orthogonality of the $\pi$ and $\pi^*(1300)$ wave functions the respective size parameters fulfil the relation $R_{\pi}=R_{\pi^*}$. The $R_\sigma$ dependence of the quarkonia decay amplitude was shown to be small \cite{mitja}. We therefore choose $R_\sigma=R_\pi$ for our numerical results presented in the forthcoming section. \newpage \section{Decay of the glueball component} \label{gluon} In this section we consider the two-body decay of the ground state scalar glueball ($G_0$) in the next-to-leading order decay mechanism of strong coupling QCD \cite{ams}. For the decay analysis we adopt a non-relativistic picture where $G_0$ is a bound state of two massive constituent gluons g. The decay is proceeding via the conversion of the constituent gluons into $Q\bar Q$ pairs as illustrated in Fig. \ref{gluea}. In the evaluation a second diagram with different quark rearrangement in the final state analogous to quarkonia decay has also to be considered. The elementary quark-gluon interaction vertex is given to lowest order in the non-relativistic limit by \begin{equation} V^{(g_i Q_l\bar Q_k)}=\gamma~ \delta(\vec q_i-\vec p_l-\vec p_k)~ (\vec{\sigma}_{(lk)}\cdot\vec{\epsilon}_{i})~{\mathbf{1}}_F^{(lk)}~ (\frac{1}{2}\sum_{a=1}^8 \lambda^a_{(lk)}A^a_i) \end{equation} with the internal momenta $\vec q_i$ for gluon i=1,2 and $\vec p_{l(k)}$ for the quarks with label l=1,3 (k=2,4). The identity operator ${\mathbf{1}}^{(lk)}_F$ projects onto a flavor singlet state of the created $Q\bar Q$ pair (lk). The last term in Eq. (15) is the color part of the interaction vertex with the Gell-Mann matrices $\lambda^a$ acting in color space of (lk). The color octet wave function of the gluon i with polarisation vector $\vec{\epsilon}_i$ is denoted by $A^a_i$. The strength of the $g_i \rightarrow Q_l\bar Q_k$ transition is given by $\gamma$. We assume that flavor symmetry is manifested in the quark-gluon coupling \cite{ams}, hence $<n\bar n|V|g> \equiv <s\bar s|V|g>=\gamma$. In order to evaluate the decay amplitude $M_{G_0 \rightarrow BC}$ we will construct a non-relativistic digluonium wave function from cavity QCD. \subsection{Digluonium wave function} In constructing the glueball wave function we use the Bag Model as a guidance. The ground state scalar glueball is realized in the Bag Model by the coupling of two transverse electric (TE) modes of the constituent gluon $(J^{PC}=1^{+-})$ to scalar quantum numbers $J^{PC}=0^{++}$ \cite{jaffe}. Using the non-relativistic TE gluon mode, proposed in Ref. \cite{iddir}, the scalar glueball wave function is in simple correspondence with the Bag Model solution and can be written as \begin{eqnarray} \Psi_{G_0}(\vec r_1,\vec r_2) &= &[TE_1(1^{+-})\otimes TE_2(1^{+-})]_{J=0} \\ & = & \left\{~ \left[ \Psi_{n_1=0l_1=1}(\vec r_1)\otimes \epsilon^\mu_1\right]_{1} \otimes \left[ \Psi_{n_2=0l_2=1}(\vec r_2)\otimes \epsilon^{\mu^\prime}_2\right]_{1}~\right\}_{J=0} \Psi_C^{(12)} \nonumber \end{eqnarray} where $\Psi_{n=0l=1}(\vec r)$ is the single particle harmonic oscillator wave functions as defined in Appendix A. The spherical components of the gluon polarisation vector $\vec \epsilon$ are given by $\epsilon^{\mu(\mu^\prime)}$. The color singlet wave function $\Psi_C^{(12)}$ for the two gluon state is \begin{equation} \Psi_C^{(12)} = \frac{1}{\sqrt{8}} \sum_{b=1}^8 A_1^b~A_2^b \end{equation} where we have chosen the "cartesian" basis $A^b$ with \begin{equation} A^b = \frac{1}{\sqrt{2}} \sum_{ij=1}^3 \left(\lambda_{ij}\right)^b c_i~\bar{c}_j \end{equation} for the octet representation of color $SU(3)$ with $c_i$ and $\bar c_j$ being the basis states for the fundamental representations ${\bf 3}$ and $\bar{{\bf 3}}$. \\ $\Psi_{G_0}(\vec r_1,\vec r_2)$ can be recoupled as \begin{eqnarray} \Psi_{G_0}(\vec r_1,\vec r_2) & = & \sqrt{3}~ \sum_{g}~ (-)^{g+1} \left\{ \begin{array}{ccc} 1 & 1& 1 \\ 1 & 1& g \end{array} \right\}~ \sqrt{2g+1} \\ & & \cdot \left\{~\left[\Psi_{n_1=0l_1=1}(\vec r_1)\otimes \Psi_{n_2=0l_2=1}(\vec r_2) \right]_g \otimes \left[\epsilon^\mu_1 \otimes \epsilon^{\mu^\prime}_2 \right]_g~\right\}_{J=0} \Psi_C^{(12)} . \nonumber \end{eqnarray} Introducing relative and centre-of-mass coordinates $\vec r=\vec r_1-\vec r_2$ and $\vec{R}=1/2(\vec{r}_1+\vec{r}_2)$, respectively, we perform the expansion \cite{mosh} \begin{eqnarray} \left[\Psi_{n_1=0l_1=1}(\vec r_1)\otimes \Psi_{n_2=0l_2=1}(\vec r_2)\right]_g & = & \sum_{nlNL}<nl,NL,g|n_1=0l_1=1,n_2=0l_2=1,g> \nonumber\\ & &~~ \left[\Psi_{nl}(\vec r)\otimes \Psi_{NL}(\vec R)\right]_g \end{eqnarray} where energy conservation restricts the sum to $2n+l+2N+L=2$. The relevant values for the transformation brackets $<nl,NL,g|n_1l_1,n_2l_2,g>$ can be found in Appendix C. \\ When projecting out the spurious centre-of-mass excitations, the normalised glueball wave function can be written in its centre-of-momentum frame as \begin{eqnarray} \Psi_{G_0}(\vec q_1,\vec q_2) & = & \left( \sqrt{\frac{5}{9}}~\left \{~ \Psi_{n=0l=2}(\vec q_1, \vec q_2)\otimes \left[ \epsilon^\mu_1 \otimes \epsilon_2^{\mu^\prime} \right]_2~ \right\}_0 \right. \\ & & \left.- \frac{2}{3}~ \left\{~\Psi_{n=1l=0}(\vec q_1, \vec q_2)\otimes \left[ \epsilon^\mu_1 \otimes \epsilon_2^{\mu^\prime} \right]_0~\right\}_0~\right) \delta(\vec q_1 + \vec q_2)~ \Psi_C^{(12)} \nonumber \end{eqnarray} with the internal gluon momenta $\vec q_i$ $(i=1,2)$ and $\Psi_{nl}(\vec q_1, \vec q_2)$ given in Appendix A. \subsection{Digluonium decay amplitudes} With the digluonium wave function of Eq. (21) were are now in the position to evaluate the decay amplitudes for the process $G_0 \rightarrow BC$ of Fig. \ref{gluea}. The decay amplitude $M_{G_0 \rightarrow BC}^{(l_{BC})}$, as defined by Eq. (3), separates in analogy to quarkonia decay into a flavor part $M^F_{G_0 \rightarrow BC}$ with \begin{equation} M^F_{G_0 \rightarrow BC}~=~ <\chi_F^B(13)\chi_F^C(42)|{\mathbf{1}}_F^{(12)} {\mathbf{1}}_F^{(43)}|0> \end{equation} and a spin-spatial-color part \begin{eqnarray} M_{G_0 \rightarrow BC}^{(l_{BC})SSC} & = & \sqrt{2\pi K}~\sqrt{\frac{E_BE_C}{M_{f_0}}}~ M_{G_0 \rightarrow BC}^C~ \int d\vec{q}_1 d\vec{q}_2 \prod_{i=1,..4}d\vec{p}_i~ \delta(\vec{q}_1-\vec{p}_1-\vec{p}_2) \\ & & \delta(\vec{q}_2-\vec{p}_4-\vec{p}_3) ~ \Psi_C^\dagger(\vec{p}_4,\vec{p}_2)~\Psi_B^\dagger(\vec{p}_1,\vec{p}_3) ~(\vec{\sigma}_{(12)}\cdot \vec{\epsilon_1})~ (\vec{\sigma}_{(43)}\cdot \vec{\epsilon_2}) ~\Psi_{G_0}(\vec{q}_1,\vec{q}_2). \nonumber \end{eqnarray} The flavor matrix elements in the $SU(3)$ flavor symmetry limit are given in Table \ref{tab2}. The color transition amplitude $M_{G_0 \rightarrow BC}^C$ is given by \begin{eqnarray} M_{G_0 \rightarrow BC}^C & = & \frac{1}{4} < \chi^{(13)}_C \chi^{(42)}_C| \sum_{c=1}^8 \lambda^c_{(12)}A^c_1~ \sum_{d=1}^8 \lambda^d_{(43)}A^d_2|\Psi_C^{(12)}> \\ & = & \frac{1}{24\sqrt{2}}~\sum_{b=1}^8 ~Tr[(\lambda^b)^2]~=~\frac{\sqrt{2}}{3} \nonumber . \end{eqnarray} The analytic evaluation of Eq. (22) contains the spin matrix element \begin{eqnarray} <\chi_{S_B}(13)~\chi_{S_C}(42)| ~(\vec{\sigma}_{(12)}\cdot \vec{\epsilon_1})~ (\vec{\sigma}_{(43)}\cdot \vec{\epsilon_2})| \left[ \epsilon^\mu_1 \otimes \epsilon_2^{\mu^\prime} \right]_{S=0,2}> & \\ =6\sqrt{(2S_B+1)(2S_C+1)}~(-)^{S_C+1}~<S_BM_BS_CM_C|S\mu+\mu^\prime> & \nonumber \left\{ \begin{array}{ccc} 1/2 & 1/2 & S_B \\ 1/2 & 1/2 & S_C \\ 1 & 1 & S \end{array} \right\} & \end{eqnarray} and the spatial overlap \newpage \begin{eqnarray} & & \int d\vec{q}_1 d\vec{q}_2 \prod_{i=1,..4}d\vec{p}_i~ \delta(\vec{q}_1-\vec{p}_1-\vec{p}_2)~ \delta(\vec{q}_2-\vec{p}_4-\vec{p}_3)~ \delta(\vec{q}_1+\vec{q}_2) \\ & & \cdot \delta(\vec{p_1}+\vec p_3 -\vec{K})~ \delta(\vec{p_4}+\vec p_2 +\vec{K})~ \Psi_{n_Bl_B}^\dagger(\vec{p}_4,\vec{p}_2)~ \Psi_{n_Cl_C}^\dagger(\vec{p}_1,\vec{p}_3)~ \Psi_{n=0,1l=2,0}(\vec{q}_1,\vec{q}_2) \nonumber \\ & & = ~\delta_{l_f;0}~\delta_{m_f;0}~\frac{1}{8}~ \int d\vec{q}_1~d\vec{\rho}~~ \Psi_{n_Bl_B}(\vec q_1-1/2(\vec K-\vec{\rho}),-\vec q_1-1/2(\vec K+\vec{\rho}))~~ \nonumber\\ & & ~\cdot~ \Psi_{n_Cl_C}(1/2(\vec K+\vec{\rho}),1/2(\vec K-\vec{\rho}))~~ \Psi_{n=0,1l=2,0}(\vec{q}_1,-\vec q_1) \nonumber \end{eqnarray} where $l_f$ is the relative orbital angular momentum of the final state with projection $m_f$. A complete listing of the analytic results for the amplitudes $M_{G_0 \rightarrow BC}^{(0)SSC}$ is given in Appendix D. \\ For the decay $G_0 \rightarrow BC~(l_f)$ we obtain the dynamical selection rule that the final state partial wave is restricted to $l_f=0$. Hence the decay $f_0(M) \rightarrow a_1\pi$ will proceed entirely by the quarkonia components of $f_0(M)$. Moreover, the D-wave decay amplitude $(l_f=2)$ for the two vector meson final state is forbidden in contrast to quarkonia decay. \\ With the $Q\bar Q$ meson size parameters fixed, $M_{G_0 \rightarrow BC}^{(0)SSC}$ depends on the glueball size parameter $R_{G_0}$. In the constituent picture we expect the mass $M_{G_0}$ of the digluonium to be $M_{G_0}\approx 2~m_g$ with a lattice motivated constituent gluon mass of about $m_g \approx 750-850~MeV$. Assuming equal confinement strength for quarks and gluons on the constituent level we get $R_{G_0}=3.65 /\sqrt{2.5}~GeV^{-1}$. With this choice the constituent mass for the lightest quark flavors lies in a reasonable mass range of $300-340~MeV$. \newpage \section{Two-body decays of scalar-isoscalar states and three-state mixing} \label{clo-wein} In the following we discuss the application of the decay models developed in the previous sections to the phenomenology of the $f_0$ decays. Given the detailed experimental decay data on the $f_0(1500)$, our analysis will focus on this state with decay predictions given for the partner states $f_0(1370)$ and $f_0(1710)$. After a brief overview of the experimental results for $f_0(1500)$, we will illustrate the distinct decay patterns resulting from the unmixed $Q\bar Q$ and $G_0$ components. In a first analysis we apply the decay models to the three-state mixing schemes constructed in Refs. \cite{ams,weina}. Then we directly extract the mixing coefficients from a fit to the experimental two-body decay data. \subsection{Observed decay properties of $f_0(1500)$} The scalar-isoscalar resonance $f_0(1500)$ is now clearly established in proton-antiproton ($p\bar p$) annihilation reactions by the Crystal Barrel collaboration at CERN \cite{amsler}. Mass and width of $f_0(1500)$ are determined as \cite{amsler}: \begin{equation} \label{1500} M_{f_0(1500)}=1505~\pm~9~MeV \qquad \Gamma_{f_0(1500)}~=~111~\pm~12~MeV . \end{equation} For the two-body decay modes $f_0(1500) \rightarrow BC$ Crystal Barrel obtains the branching ratios $BR(BC)$: \begin{eqnarray} BR(\pi\pi) = 29\pm7.5 \% & ~~BR(\eta\eta) = 4.6\pm1.3 \% & ~~BR(\eta\eta^\prime) = 1.2\pm0.3\% \nonumber\\ BR(K\bar K)= 3.5\pm0.3\% & ~~BR(4\pi) = 61.7\pm9.6\%& . \end{eqnarray} Here, the $4\pi$ final state includes $\sigma\sigma$ and $\pi^*(1300)\pi$ as intermediate decay channels. The allowed $\rho\rho$ decay mode of the $f_0(1500)$ is found to be weak \cite{thoma}. Taking the central value for the width we deduce from the measured branching ratios the partial decay widths $\Gamma(BC)$: \begin{eqnarray} \label{expcrystal} \Gamma(2\pi) = 32.2\pm8.4~MeV & ~~\Gamma(\eta\eta) = 5.2\pm1.5~MeV & ~~\Gamma(\eta\eta^\prime) = 1.4\pm0.4~MeV \nonumber\\ \Gamma(K\bar K) = 3.9\pm0.4~ MeV & ~~\Gamma(4\pi) = 68.5\pm10.7~MeV & . \end{eqnarray} Alternative multi-channel analyses \cite{bugga} based on data sets from previous experiments have also been performed. The results for the $K \bar K$ and $\eta\eta$ modes are within the experimental errors of the Crystal Barrel data. A larger total width ($\Gamma_{tot}=132\pm15~MeV$) and $\pi\pi$ decay width ($\Gamma_{\pi\pi}=60\pm12~MeV$) is obtained suggesting a smaller value for $\Gamma(4\pi)$ than indicated in Eq. (28). \subsection{Two-body decays of the $Q\bar Q$ and $G_0$ components} We first discuss the decay patterns resulting from the individual decay mechanisms of Figs. \ref{quarkonia} and \ref{gluea}, coupling to the $Q\bar Q$ and the $G_0$ components of the $f_0$ states, respectively. \\ In lowest order, neglecting a possible $G_0 \rightarrow G_0G_0$ contribution, the decay $f_0 \rightarrow BC$ will proceed via its quarkonia components $Q\bar Q$. In general we consider a mixed quarkonium state \begin{equation} \label{mixq} |Q\bar Q>~=~cos\alpha~|n\bar n>~-~sin\alpha~|s\bar s>. \end{equation} The dependence of the branching ratios $BR$ for the two-body decays of $f_0(1500)$ on the mixing angle $\alpha$ is given in Fig. \ref{loword}. The results of Fig. \ref{loword} correspond to the original discussion in Ref. \cite{ams} for the two pseudoscalar meson decay modes of a $^3P_0$ isoscalar $Q\bar Q$ state. But now predictions for all possible two-body decay modes are given. \\ With the three-state mixing schemes of Ref. \cite{ams} ($a_{n\bar n}=0.43, a_{s\bar s}=-0.61$, corresponding to $\alpha=54.8^o)$ and Ref. \cite{weina} $(a_{n\bar n}=0.40,a_{s\bar s}=-0.90,$ that is $\alpha=66^o)$ we obtain following ratios for the pseudoscalar decay modes \begin{eqnarray} BR(\pi\pi):BR(K\bar K): BR(\eta\eta) : BR(\eta\eta^\prime)& & \nonumber\\ =~1: 0.21 : 0.005 : 0.93 \qquad \mbox{Ref. \cite{ams}} & & \\ =~ 1 : 0.97 : 0.12 : 1.80 \qquad \mbox{Ref. \cite{weina}} & & \end{eqnarray} to be compared with the Crystal Barrel results \cite{amsler} of \begin{eqnarray} BR(\pi\pi):BR(K\bar K): BR(\eta\eta) : BR(\eta\eta^\prime) & & \\ =~~ 1: 0.12 \pm 0.03 : 0.16 \pm 0.06 : 0.04 \pm 0.02 . & & \nonumber \end{eqnarray} Best agreement with data is achieved for $f_0(1500) \approx |n \bar n>$, for which we obtain \begin{equation} BR(\pi\pi):BR(K\bar K): BR(\eta\eta) : BR(\eta\eta^\prime) =1 : 0.19 : 0.16 : 0.10 . \end{equation} Hence, experimental data for the ratios of two pseudoscalar decay modes of the $f_0(1500)$ are consistent with a dominant $n\bar n$ interpretation. However, the predicted total width for a pure $^3P_0~n\bar n$ state is in conflict with experimental observation. In Fig. \ref{width} we indicate the calculated total width for a $f_0(1500)$ state with the $Q\bar Q$ configuration of Eq. (\ref{mixq}). A pure $n\bar n$ state results in $\Gamma_{tot}\approx500~MeV$, which is considerably higher than the observed width. A sizable reduction of the $Q\bar Q$ amplitude in the $f_0(1500)$, as affected by the admixture of a $G_0$ component, leads to a reduction of the total width. For example, the mixing schemes of Refs. \cite{ams} and \cite{weina} lead to a total width of $\Gamma=111~MeV$ and $\Gamma=127~MeV$, respectively. Working in the lowest order of the decay mechanism, i.e. $f_0(1500)$ decays by its $Q\bar Q$ components, the observed two pseudoscalar decay modes are consistent with a dominant $n\bar n$ content, slightly favoring the mixing scheme of Ref. \cite{ams}; the possible presence of an admixed $G_0$ component is then only reflected in the corresponding reduction of the total width. \\ When going beyond the two pseudoscalar meson decay modes in the leading order decay scheme, strong deviations from the observed decay pattern of the $f_0(1500)$ arise \cite{mitja}. For the two-body decay modes leading to the $4\pi$ decay channel, including also charged pion combinations, we obtain \begin{equation} \label{xy} BR(\pi\pi):BR(\rho\rho):BR(\pi^*\pi):BR(\sigma\sigma):BR(a_1\pi)= 1: 0.97 :0.14 :0.31 :1.8 . \end{equation} Predictions are independent of the mixing angle $\alpha$, since the final states of Eq. (\ref{xy}) only couple to the $n\bar n$ configuration. Relative to the $\pi\pi$ channel the $a_1\pi$ mode is the strongest decay channel. The predicted large $\rho\rho$ branching ratio with $BR(\rho\rho)\approx BR(\pi\pi)$ is in conflict with the preliminary analysis \cite{thoma}, where the $\rho\rho$ channel is found to be strongly suppressed. Furthermore, for the combined $\sigma\sigma$ and $\pi^\ast\pi$ decay channels with $BR(\pi^\ast \rightarrow \sigma\pi)=0.1$ \cite{ulrike} we deduce the ratio $BR(4\pi)/BR(2\pi)=0.32$, which is much lower than the experimental result of $BR(4\pi)/BR(2\pi)= 2.1 \pm 0.6$ \cite{amsler}. The observed suppression of the $\rho\rho$ and enhancement of the $4\pi^0$ decay modes are both in strong conflict with a na\"{\i}ve $Q\bar Q$ interpretation of the $f_0(1500)$ and give further indication for a sizable direct coupling of the gluonic component to the decay channel.\\ Next we consider the decay pattern evolving from the direct decay of the $G_0$ component as indicated in Fig. \ref{gluea}. Branching ratios for the scalar glueball $(G_0)$ decay in dependence on its mass are shown in Fig. \ref{gluemass}. A main signature is the strong suppression of the two vector meson decay channels, particularly the $\rho\rho$, over the entire mass range, due to the dynamically forbidden D-wave amplitude. The $\pi^\ast\pi$ and $\sigma\sigma$ decay modes exceed the contribution of the $\pi\pi$ decay channel for a $G_0$ mass of about $1500~MeV$. These results are in clear contrast to those obtained for the $Q\bar Q$ decay mechanism. Both, the suppression of the $\rho\rho$ decay mode and the enhancement of the $4\pi^0$ decay channels resulting from the decay of the gluonic component $G_0$ are features observed in the $f_0(1500)$ decay pattern as discussed previously. For the two pseudoscalar meson decays the results differ from the na\"{\i}ve expectation \cite{ams} \begin{equation} \label{gl} BR(\pi\pi):~BR(K\bar K):~BR(\eta\eta): BR(\eta\eta^\prime)~=~1:~\frac{4}{3} :~\frac{1}{3}:~0 \end{equation} deduced from flavor symmetric couplings. For a glueball mass of $M_{G_0}=1500~(1700)~MeV$ we obtain \begin{equation} BR(\pi\pi):~BR(K\bar K):~BR(\eta\eta)=1:~3.7~(4):~3.9~(4.3). \end{equation} with $BR(\eta\eta^\prime)=0$ in the SU(3) flavor symmetry limit. The enhancement of the $K\bar K/\pi\pi$ decay ratio relative to the na\"{\i}ve SU(3) flavor estimate of Eq. (35) is a feature confirmed by recent lattice results \cite{sexton}. For the given choice of radial meson parameter ($R_K=2.82~GeV^{-1}, R_\eta=2.48~GeV^{-1}$), as extracted from tensor meson decay, we obtain $BR(K\bar K)\approx BR(\eta\eta)$. Choosing the SU(3) flavor limit of $R_K=R_\eta$ leads to $BR(K\bar K)=4.2~BR(\eta\eta)$, which is again consistent with the analysis of Ref. \cite{sexton}. Hence, the proposed modelling of the $G_0$ decay mechanism contains the main features obtained from lattice calculations.\\ As a measure of flavor symmetry breaking we introduce the parameter $S$ defined by \\ $S\equiv <u\bar u|V|g>/<s\bar s|V|g>\not= 1$ leading to $BR(\eta\eta^\prime)\not= 0$. The flavor matrix elements for $K\bar K, \eta\eta, \eta\eta^\prime$ decays listed in Table \ref{tab1} correspond to the exact flavor symmetry limit $(S=1)$ and have to be generalised for pure glueball decays $(a_{G_0}=1)$ \begin{eqnarray} M^F_{G_0 \rightarrow K\bar K}&=& 2~\sqrt{2}~S \nonumber\\ M^F_{G_0 \rightarrow \eta\eta}&=& \sqrt{2}~(\alpha_{PS}^2 +S^2\beta_{PS}^2) \\ M^F_{G_0 \rightarrow \eta\eta^\prime}&=& 2~\alpha_{PS}~\beta_{PS}~ (1-S^2 ) . \nonumber \end{eqnarray} If we adopt the extreme value of $S^2=0.5$ \cite{anisovich} the strength of the $K\bar K$ and $\eta\eta$ decay modes will be suppressed by 50\% and 34\%, respectively, as compared to the unbroken flavor symmetry couplings.\\ For scalar glueball masses at $M_{G_0}= 1300,~ 1500,~ 1700~MeV$ the reduced available phase space suppresses the $\eta\eta^\prime$ mode: \begin{equation} \begin{array}{cccccc} M_{G_0}~[MeV] & & BR(\pi\pi):& BR(K\bar K):& BR(\eta\eta) :& BR (\eta\eta^\prime) \nonumber\\ 1300 & & 1:& 1.3:& 2:& 0.0 \nonumber\\ 1500 & & 1:&1.8:&2.6:&0.08 \\ 1700 & & 1:& 2:& 2.9:& 0.28 \nonumber . \end{array} \end{equation} Strong violation of flavor symmetry reduces the $K\bar K/\pi\pi$ ratio, while leading to a finite $\eta\eta^\prime$ contribution. However, data on charmonium decay involving hard gluons indicate no significant symmetry breaking \cite{ams}. We therefore choose to set S=1 in the following discussion. \\ Decay data on $f_0(1500)$ possess qualitative features arising both from the decay of a $Q\bar Q$ (here dominantly $n\bar n$) and a $G_0$ configuration, where a pure $Q\bar Q$ or $G_0$ state interpretation of the $f_0(1500)$ is excluded. Given this simple first analysis of the decay pattern of the $f_0(1500)$ we are naturally led to a mixing scheme where both components are present. \subsection{Decay analysis in the three-state mixing schemes} In the following we give a decay analysis for the $f_0$ states defined by the three-state mixing schemes of Refs. \cite{ams} and \cite{weina}. Now, the couplings both of the $Q\bar Q$ and $G_0$ components to the two-meson decay channels are included. \\ When the physical $f_0$ mesons are given as eigenstates of the interaction Hamiltonian $H_I$ of Eq. (1), the partial decay widths of the process $f_0 \rightarrow BC$ are defined by Eq. (5) with the ratio of coupling strengths $\kappa$ and the phase angle $\phi$ left as free parameters. The structure of $H_I$ implies unbroken flavor symmetry, hence we choose $S=1$. For the $f_0(1500)$ decays we give in Table \ref{tab3} the squared decay amplitudes and the interference terms as defined in Eq. (5), where mass averaging as given by Eq. (6) is already included. The mass distribution of $f_0(1500)$ is parameterised with the values of Eq. (\ref{1500}). \\ The dependence of the partial decay widths $\Gamma(BC)$ on the mixing coefficients $a_i$ $(i=n\bar n, s\bar s, G_0)$ in the schemes of Ref. \cite{ams} ($\equiv$ scheme A) and Ref. \cite{weina} ($\equiv$ scheme B) are studied by adjusting $\kappa$ and $\phi$ to the Crystal Barrel data for $f_0(1500) $ decays. First we have determined $\kappa$ and $cos\phi$ from a fit to the observed decay modes $f_0(1500) \rightarrow K\bar K$ and $\pi\pi$. For both mixing schemes we obtain parameter values resulting in a $ f_0 \rightarrow \eta\eta$ decay width in good agreement with the experimental number. This implies phenomenologically, that a $G_0 \rightarrow G_0 G_0$ contribution to the $\eta\eta$ decay channel can be neglected. For this reason, we also take the $\eta\eta$ decay mode into account when determining $\cos\phi$ and $\kappa$ from a more restrictive fit. The partial decay widths have a rather weak dependence on the parameterisation of the mass distribution of the physical $f_0$ states. Therefore, the discussion \cite{amsler} on the detailed relationship between the $f_0(1370)$ and $f_0(400-1200)$ states is peripheral to our analysis. For our numerical results we use the resonance parameters $M_{f_0(1370)}=1360~MeV,~ \Gamma_{f_0(1370)}=350~MeV$ \cite{amsler}, and $M_{f_0(1710)}=1697~MeV,~ \Gamma_{f_0(1710)}=175~MeV$ \cite{pdg}.\\ The mixing coefficients in scheme A \cite{ams} are given as \begin{equation} \left( \begin{array}{c} |f_0(1370)> \\ |f_0(1500)>\\ |f_0(1710)> \end{array} \right) = \left( \begin{array}{ccc} 0.86 & 0.13 & -0.5 \\ 0.43 & -0.61 & 0.61 \\ 0.22 & 0.76 & 0.6 \end{array} \right)~ \left( \begin{array}{c} |n\bar n> \\ |s\bar s> \\ |G_0> \end{array} \right). \end{equation} From a fit to the $f_0(1500) \rightarrow \pi\pi, K\bar K, \eta\eta$ decay widths we deduce $cos\phi=-0.67$ and $ \kappa=0.095 ~GeV^{-2}$. Predictions for the partial decay widths of the physical $f_0$ states are presented in Table \ref{tab4}. For $f_0(1500)$ the predicted $\eta\eta^\prime$ decay width is considerably lower than the Crystal Barrel value, but the obtained ratio $\Gamma(\eta\eta^\prime)/\Gamma(\eta\eta)$ agrees with the GAMS value of $2.7\pm0.8$ for $f_0(1590)$ \cite{pdg}. It is generally believed that $f_0(1500)$ and $f_0(1590)$ are manifestations of a single state, but then the large deviation in the experimental data for the $\eta\eta^\prime$ decay mode has to be clarified. Compared to the lowest order decay analysis, where only the coupling to the $Q\bar Q$ component is retained, the $\rho\rho$ mode is more suppressed and the $4\pi$ channel enhanced. We obtain $BR(\pi\pi):BR(\rho\rho):BR(4\pi)= 1:~0.65:~0.46-0.76$ for $BR(\pi^* \rightarrow \sigma\pi)=0.1 - 1$, still in deviation from the Crystal Barrel data. \\ In the mixing scheme B \cite{weina} the amplitudes for the physical states are given as \begin{equation} \left( \begin{array}{c} |f_0(1370)> \\ |f_0(1500)>\\ |f_0(1710)> \end{array} \right) = \left( \begin{array}{ccc} 0.84& 0.28 & -0.46\\ 0.40 & -0.9 & 0.19 \\ 0.36 & 0.34 & 0.87 \end{array} \right)~ \left( \begin{array}{c} |n\bar n> \\ |s\bar s> \\ |G_0> \end{array} \right). \end{equation} From an analogous fit we obtain $cos\phi=-0.953$ and $\kappa=0.437~GeV^{-2}$. The predicted two-body partial decay widths are given in Table \ref{tab5}. Compared to the mixing scheme A, $f_0(1500)$ has a reduced $G_0$ component, but the experimental data demand a significant $G_0 \rightarrow BC$ contribution in fitting the $\pi\pi,~K\bar K$ and $\eta\eta$ decays. This is balanced by giving the $G_0 \rightarrow BC$ transition a stronger weight than in the mixing scheme A, resulting in a higher value for $\kappa$. Again, the predicted strength of the $\eta\eta^\prime$ decay mode overestimates the experimental value, while similar results are obtained for the $\rho\rho$ and $4\pi$ decay modes. For mixing scheme B we obtain $BR(\pi\pi):BR(\rho\rho):BR(4\pi)= 1:~0.46:~0.55-0.97$ for $BR(\pi^* \rightarrow \sigma\pi)=0.1 - 1$. In both schemes $a_1\pi$ is a dominant decay mode with $BR(\pi\pi)\approx BR(a_1\pi)$. With our fitting procedure, that is adjusting the relative strength of the $G_0$ decay contribution to the $\pi\pi, K\bar K$ and $\eta\eta$ decay channels of the $f_0(1500)$, similar results are obtained for the decay pattern of the $f_0(1500)$ in both mixing schemes. Particular predictions for the decay modes of the partner states $f_0(1370)$ and $f_0(1710)$ can be dramatically different in the two mixing schemes. In scheme A the total width for $f_0(1710)$ is in good agreement with the experimental result of $\Gamma=175\pm9~MeV$ $(\Gamma=133\pm14~MeV)$ \cite{pdg}, but scheme B predicts a unobservable decay width of $\Gamma \approx 2000~MeV$. Furthermore, for $f_0(1710)$ we get \begin{equation} BR(\pi\pi)/B(K\bar K) )=\left\{ \begin{array}{ccc} 0.93 & \mbox{for} & \mbox{scheme A} \\ 0.60 & \mbox{for} & \mbox{scheme B} \end{array} \right. \end{equation} to be compared with the experimental value of $0.39\pm0.14$ \cite{pdg}.\\ The two mixing schemes predict rather similar results for the $f_0(1370)$ state. In both schemes $f_0(1370)$ has dominant $\rho\rho$ and $a_1\pi$ decay modes. The ratio $BR(K\bar K)/(BR(\eta\eta)$ is found to be \begin{equation} BR(K\bar K)/BR(\eta\eta)=\left\{ \begin{array}{ccc} 2.6 & \mbox{for} & \mbox{scheme A} \\ 1.45 & \mbox{for} & \mbox{scheme B} \end{array} \right. \end{equation} to be compared with the Crystal Barrel value of $ BR(K\bar K)/BR(\eta\eta) \approx 2.5$ \cite{amsler}. \\ The main conclusions of this section remain unchanged if we use for $f_0(1500)$ a larger $\pi\pi$ decay width as reported in Ref. \cite{bugga}, since a corresponding fit leads to similar values for $\kappa$ and $cos\phi$ as for the adjustment to the Crystal Barrel data. \\ Inclusion of the direct coupling of the $G_0$ component of the mixed $f_0(1500)$ state to the two-meson decay channels gives an improved description of its decay features. For both proposed mixing schemes predictions for the $f_0(1500)$ decays are very similar, since the difference in size of the $G_0$ amplitude is compensated by adjusting the strength of the $G_0 \rightarrow (Q\bar Q)~(Q\bar Q)$ transition. \\ Also, predictions for the $f_0(1370)$ state do not differ too much, since both schemes assign a dominant $n\bar n$ component to this state. The key difference rests on predictions for the partner state $f_0(1710)$. We have shown that a reduced $G_0$ component in $f_0(1500)$, as proposed by Ref. \cite{weina}, is incompatible with the existence of a $f_0$ state at 1700 MeV as a narrow resonance. Conversely, if the existence of a narrow scalar component of the $f_{j=0,2}(1710)$ \cite{pdg} state is confirmed, then a non-negligible glueball component is residing in the $f_0(1500)$ state. \newpage \subsection{Mixing amplitudes} \label{analysis} Given the shortcomings of the proposed mixing schemes of Refs. \cite{ams,weina} we now directly extract the mixing amplitudes of the $f_0$ states in the given decay formalism. We proceed in an analogous fashion, requiring a fit to the experimental data on $f_0(1500)$ decays, while demanding the existence of a narrow resonance $f_0(1710)$ as an additional requirement.\\ First, we start with the unspecified mixing amplitudes of the $f_0(1500)$ state vector as defined by Eq. (2). The mixing amplitudes of $f_0(1500)$ are determined from a fit to the experimental data of the $\pi\pi$, $K\bar K$ and $\eta\eta$ decay modes with the restriction of a weak $\rho\rho$ decay channel. Exact flavor symmetry ($S=1$) for the quark-gluon coupling is assumed. Good agreement with the experimental data requires $a_{n\bar n}\cdot a_{s\bar s}<0$, i.e. a destructive interference between the $n\bar n$ and $s\bar s$ flavors, and we get \begin{equation} |f_0(1500)>~=~0.314~|n \bar n>~-~0.581~|s\bar s>~+~0.751~|G_0> \end{equation} with the ratio of coupling strengths $\kappa=0.1~GeV^{-2}$ and the phase factor $cos\phi=-0.92$. The fit to the $f_0(1500)$ decays alone does not fix the phase of the $a_{G_0}$ amplitude relative to the quarkonia coefficients. Instead we obtain the condition that $a_{G_0}\cdot cos\phi<0$. For the $f_0(1500)$ state vector we have taken $a_{G_0}>0$ and $cos\phi<0$ for reasons given later. The relative phases between the state amplitudes of Eq. (43) are the same as in mixing schemes A and B. For the full set of partial decay widths we obtain the results given in Table \ref{tab6}. Compared to mixing scheme A the $\rho\rho$ mode of $f_0(1500)$ is suppressed by a factor of two. This is achieved by enhancing the $G_0$ amplitude since the $\rho\rho$ suppression for $f_0(1500)$ is driven by the gluonic component (see Sec. \ref{gluon}). An alternative fit to obtain $\rho\rho$ suppression requires a strong $s\bar s$ component and a larger value for $\kappa$ as resulting for example from the mixing scheme B. However, the existence of a narrow $f_0(1710)$ state rules out this possibility. \\ For the $4\pi/2\pi$ decay branching ratio we obtain $BR(4\pi)/BR(2\pi)=0.57-1.04$ using $BR(\pi^* \rightarrow \sigma\pi)=0.1-1$, where recent results \cite{ulrike} for $ \pi^\ast \rightarrow \sigma\pi$ favor the lower value. This is still in deviation from the Crystal Barrel value of $BR(4\pi)/BR(2\pi)=2.1\pm0.6$ \cite{amsler}. The disagreement possibly hints at a sizable $G_0$ component in $\sigma$. A strong coupling of $\sigma$ to $G_0$ is supported by the experimental finding \cite{pdg} that $\sigma\sigma$ is the largest hadronic branching ratio for the charmonium states $\chi_0$ and $\chi_2$. In this case, the lowest order decay mechanism $G_0 \rightarrow G_0 G_0$, neglected here, can enhance the $\sigma\sigma$ and therefore the $4\pi$ decay mode. The predicted ratio $BR(\eta\eta^\prime)/BR(\eta\eta)=2.2$ is in agreement with the GAMS value of $2.7\pm0.8$ \cite{pdg}, but in conflict with the Crystal Barrel result of $0.27\pm0.10$ \cite{amsler}. Breaking of flavor symmetry at the quark-gluon vertex, that is $S \not= 1$, will further enhance the $\eta\eta^\prime$ decay width. \\ With mass and eigenstate of $f_0(1500)$ fixed at $1500~MeV$ and by Eq. (43), respectively, we deduce for the bare masses $m_i$ $(i=n\bar n,s\bar s, G_0)$ entering in $H_I$ of Eq. (1) \begin{eqnarray} m_{n\bar n}=1500~MeV-3.38~z,~ & ~~m_{s\bar s}=1500~MeV+1.29~z, \\ m_{G_0}=1500~MeV+0.18~z & \nonumber \end{eqnarray} with the mixing strength $z>0$. For the bare masses we obtain the level ordering $m_{n\bar n}<m_{G_0}<m_{s\bar s}$, which is independent of z and corresponds to that proposed originally in Ref. \cite{ams}. When taking the alternative sign pattern of $a_{G_0}<0$ and $cos\phi>0$ for the fit to the $f_0(1500)$ decays, we obtain the unphysical level ordering of the bare states with $m_{n\bar n}>m_{s\bar s}>m_{G_0}$, we hence exclude this choice. Additionally, for the masses $M_{<(>)}$ of the mixing partners lying above $(M_>)$ and below $(M_<)$ $f_0(1500)$ we obtain: \begin{eqnarray} M_> & = & 1500~MeV~+2~z \\ M_< & = & 1500~MeV~-3.9~z. \nonumber \end{eqnarray} For consistency, we call these states $f_0(1370)$ and $f_0(1710)$ and use the values for mass and total width given in the previous section for the parameterisation of the respective mass distributions. For the mixing partners of the $f_0(1500)$ we obtain the glueball-quarkonia content \begin{equation} |f_0(1710)>~=~0.149~|n\bar n>~+~0.811~|s\bar s>~+~0.565~|G_0> \end{equation} and \begin{equation} |f_0(1370)>~=~0.938~|n\bar n>~+~0.070~|s\bar s>~-~0.341~|G_0>, \end{equation} where the amplitudes are independent of the particular value of z. Predictions for the corresponding partial decay widths of $f_0(1710)$ and $f_0(1370)$ are also given in Table \ref{tab6}. The total width and the ratio $BR(\pi\pi)/BR(K\bar K)=0.57$ of the $f_0(1710)$ are in good agreement with the experimental results of of $\Gamma=175\pm9~MeV$ $(133\pm14~MeV)$ \cite{pdg} and $BR(\pi\pi)/BR(K\bar K)=0.39\pm0.14$ for $f_{j=0,2}(1710)$ \cite{pdg}. For the $f_0(1370)$ the predicted ratio of $BR(K\bar K)/BR(\eta\eta)=2.7$ is to be compared to the experimental value $\approx 2.5$ \cite{amsler}. For the total width we get $\Gamma_{tot}=479~MeV$, somewhat larger than the Crystal Barrel value of $\Gamma=351\pm41~MeV$ \cite{amsler}. \\ For a mixing energy $z=43\pm31~MeV$ \cite{weinb}, as deduced from lattice QCD in the quenched approximation, we obtain $m_{G_0}=1508\pm6~MeV$, in agreement with the lattice result of $m_{G_0}=1.65\pm0.15~GeV$ \cite{tep}. For the bare masses we get $m_{n\bar n}=1355\pm105~MeV$ and $M_{s\bar s}=1556\pm40~MeV$. The values for the glueball-quarkonia mixing matrix element z favored by previous approaches are $z=77~MeV$ \cite{weina}, $z=64\pm13~MeV$ \cite{weinb} and $z\approx 100~MeV$, where later value is deduced from the full nonperturbative three-state mixing scheme of Ref. \cite{ams}. Hence, semiphenomenological extractions of z favor a value near the upper limit set by the lattice result. \\ Choosing an intermediate value of $z\approx 80~MeV$ we obtain for the physical masses of the mixing partners given by Eq. (45), $M_>=1660~MeV$ and $M_<=1190~MeV$, where $M_<$ is lower than the mass of the $f_0(1370)$. A similar result is obtained from the full mixing approach of Ref. \cite{ams}, which corresponds qualitatively to our extracted scheme. \\ Given our fit to the $f_0(1500)$ decays, we obtain a physical mass $M_<$ and a total width which deviate from the corresponding data on $f_0(1370)$. Therefore, within the framework of a three-state mixing scheme, the lower lying partner of $f_0(1500)$ could be identified with the very broad structure around 1100 MeV \cite{ams}, called $f_0(400-1200)$ \cite{pdg}. Alternatively, the Crystal Barrel state $f_0(1370)$ could be the high mass tail of $f_0(400-1200)$. However, given the ansatz of Eq. (1), the most stable and testable consequences of our mixing scheme concern predictions for all the two-body decay modes of the $f_0$ states as given in Table \ref{tab6}. Given the sensitivity of the decay pattern on the particular size of the $G_0$ and $Q\bar Q$ components in the $f_0$ states only, a full experimental determination provides a stringent test. \newpage \section{Summary and conclusions} \label{summary} We have performed a detailed study of the two-body decay properties of the scalar-isoscalar $f_0(1370),~f_0(1500)$ and $f_0(1710)$ states resulting from the mixture of the lowest lying scalar glueball with the isoscalar quarkonia states of the $0^{++}$ nonet. In the decay analysis we have taken into account the coupling of the quarkonia and glueball components of the $f_0$ states to the quarkonia components of the two-meson final state, where the decay dynamics is guided by strong coupling QCD. Leading order corresponds to the transitions $Q\bar Q \rightarrow (Q\bar Q)~(Q\bar Q)$ and $G_0 \rightarrow G_0~G_0$, where the latter transition, which can contribute to $\eta\eta,~\eta\eta^\prime$ and $\sigma\sigma$ final states, is omitted due to its ill-constrained nature. The coupling of the quarkonia component $(Q\bar Q)$ of the $f_0$ states is described in the framework of the $^3P_0$ pair creation model, which is successful in the phenomenology of OZI-allowed meson decay. In next-to-leading order we obtain a direct coupling of the $G_0$ component of the $f_0$ states to the quarkonia component of the decay channel. The corresponding transition $G_0 \rightarrow (Q\bar Q)~(Q\bar Q)$ is modelled by resorting to a scalar digluonium wavefunction for $G_0$ as given by cavity QCD. Predictions for the two-pseudoscalar decay channels, that is $\pi\pi,~K\bar K$ and $\eta\eta$, are in rough agreement with recent lattice results in the limit of unbroken flavor SU(3). \\ Our analysis on the two-body decay modes of the $f_0$ states focuses on the $f_0(1500)$, where detailed experimental data are available. In comparison to the observed decay properties of $f_0(1500)$ we elaborate the distinct decay patterns resulting from the unmixed $Q\bar Q$ and $G_0$ components. Taking into account the coupling of the $Q\bar Q$ components of the $f_0(1500)$ only, which corresponds to the leading order decay mechanism of Ref. \cite{ams}, data on two-pseudoscalar decay channels are consistent with a dominant $n\bar n$ structure residing in this state. However, $f_0(1500)$ is too narrow for a simple $n\bar n$ structure, implying a sizable reduction of its quarkonia components as resulting from glueball-quarkonia mixing. When going beyond the two-pseudoscalar decay channels large deviations from the observed decay pattern of the $f_0(1500)$ arise. The large $4\pi$ branching ratio, fed by the $\sigma\sigma$ and $\pi^\ast \pi$ decay channels, and the suppression of $\rho\rho$ is in clear conflict with the simple leading order analysis of $f_0(1500)$, where coupling to its $Q\bar Q$ component is taken, and points towards the relevance of a possible glueball component. Coupling of the $G_0$ component of the $f_0(1500)$ to the quarkonia component of the two-body final state yields a significantly different decay pattern. In the two-pseudoscalar sector strong $K\bar K$ and $\eta\eta$ decay modes are found. Furthermore, $\pi^\ast \pi$ and $\sigma\sigma$ dominate over the $\pi\pi$ decay channel, whereas $\rho\rho$ is strongly suppressed. This latter feature of a strongly reduced $\rho\rho$ coupling to $G_0$ is also obtained in a recent work \cite{jin}, where the coupling of $G_0$ to the pionic decay channel is set up in an effective linear sigma model. Both observed features of the $f_0(1500)$, the enhancement of the $4\pi^0$ decay channels and the reduction of the $\rho\rho$ channel, are driven by its gluonic component.\\ In a next step we have investigated the decay properties of the $f_0(1370), ~ f_0(1500)$ and $f_0(1710)$ states in the proposed three-state mixing schemes of Refs. \cite{ams,weina}. Due to the inclusion of the $G_0 \rightarrow (Q\bar Q)~(Q\bar Q)$ mechanism in the decay analysis we were able to discriminate between the mixing schemes, in particular with respect to the orthogonal physical consequences for the $f_0(1710)$ state. A negligible $G_0$ component in $f_0(1500)$ as proposed by Ref. \cite{weina} is incompatible with the existence of a narrow $f_0$ state in the 1700 MeV mass region; hence our first analysis of the two-body decays favors the mixing scheme of Ref. \cite{ams}. \\ Finally, we have extracted a three-state mixing scheme based on a detailed fit to the experimental $f_0(1500)$ decays, while demanding the existence of a narrow $f_0$ mixing partner in the 1700 MeV mass region. For the bare masses before mixing we obtain the level ordering $m_{n\bar n}<m_{s\bar s}<m_{G_0}$, corresponding qualitatively to the scheme originally proposed in Ref. \cite{ams}. Adjustment of the gluonic decay mechanism leads to a proper description of the observed two-pseudoscalar decay modes of $f_0(1500)$ with the exception of $\eta\eta^\prime$. Accordingly, we obtain a significant $\rho\rho$ suppression coupled at the same time to the enhancement of the $4\pi^0/2\pi$ ratio, in line with experimental requirements. A full reduction of the observed $4\pi/2\pi$ ratio requires a strong $\sigma\sigma$ mode. Both deviations in the $\eta\eta^\prime$ and $\sigma\sigma$ decay channels suggest a non-negligible leading order $G_0 \rightarrow G_0~G_0$ contribution, which was neglected here. A key signature for the extracted state composition of $f_0(1500)$ is the prediction for the $a_1\pi$ decay mode with $BR(a_1\pi) \approx BR(\pi\pi)$. A pure $Q\bar Q$ configuration or equivalently, application of the leading order decay mechanism to the mixed $f_0(1500)$ state results in $BR(a_1\pi) \approx~2 BR(\pi\pi)$. Hence an experimental determination of the $a_1\pi$ decay channel of $f_0(1500)$ is desirable to further quantify the relevance of its gluonic component. \\ The resulting predictions for the decays of the $f_0(1370)$ and $f_0(1710)$ states are consistent with the limited experimental data \cite{amsler,pdg} on pseudoscalar decay modes. Clearly a search for the decay channels feeding the $4\pi$ decay products is most relevant in clarifying the nature of these partner states. Predictions for the total widths of $f_0(1370)$ and $f_0(1710)$ roughly agree with the observed values \cite{amsler,pdg}; the physical mass spectrum resulting from the extracted mixing scheme points towards a sizable influence of the broad $f_0$ structure at 1100 MeV \cite{amsler}, which should be clarified. \\ The quantitative predictions of our analysis emphasise the role of the gluonic components in the $f_0$ states and its observable consequences in the two-body decay modes. Hence a full determination of the hadronic decay properties provides a sensitive test on the nature of the $f_0$ states, revealing the intrusion of the glueball ground state in the scalar-isoscalar meson spectrum. \begin{acknowledgments} This work was supported in part by the Graduiertenkolleg "Struktur und Wechselwirkung von Hadronen und Kernen" (DFG Mu705/3) and by a grant of the Deutsches Bundesministerium f\"ur Bildung und Forschung (contract No. 06 T\"u 887). \end{acknowledgments} \newpage \begin{appendix} \section{Spatial wave functions} For the spatial part of the wave functions used in the text we use harmonic oscillators $\Psi_{nl}(\vec p_i, \vec p_j)$ which are defined in the limit of equal constituent masses as \begin{eqnarray} \Psi_{nl}(\vec{p_i},\vec{p_j})&=& N_{nl}~\left(\frac{R}{2}\right)^l~ |\vec{p_i}-\vec{p_j}|^l~ \exp\left\{ - \frac{1}{8} R^2 ( {\vec{p}}_i-{\vec{p}}_j)^2 \right\} \nonumber\\ & & L_n^{l+1/2}(\frac{R^2}{4}(\vec{p_i}-\vec{p}_j)^2)~Y_{lm}(\hat{\vec p_i-\vec p_j}) \end{eqnarray} with the internal momentum $p_{i(j)}$ and the size parameter $R$. The normalisation constant $N_{nl}$ depends on the radial and orbital quantum numbers n, l with \begin{equation} N_{nl}=\left[~ \frac{2(n!)R^3}{\Gamma(n+l+3/2)}\right]^{1/2}. \end{equation} The Laguerre polynomials are given by \begin{equation} L_n^{l+1/2}(p)=\sum_{k=0}^{n}\frac{(-)^k~ \Gamma(n+l+3/2)}{k!~\Gamma(k+l+3/2)}~p^k. \end{equation} For the $ Q\bar Q$ mesons we use: \begin{equation} (nl) = \left\{ \begin{array}{ccl} (00)&\mbox{for}&\pi,~\eta,~\eta^\prime,K,~\bar K,~\rho,~\omega \\ (01)&\mbox{for}&a_1,~\sigma, ~f_0 \\ (10)&\mbox{for}&\pi^*(1300) \end{array}\right. \end{equation} \newpage \section{Quarkonia decay amplitudes} In the following we give the full expressions for the spin-spatial-color amplitudes $M^{(l_{BC})SSC}_{Q\bar Q \rightarrow BC}$ defined in Eq. (12) as obtained for the quarkonia decay process $Q\bar Q \rightarrow BC$. For equal size parameters ($\equiv R_B$) of the final state mesons B and C we obtain: \\ 1) $Q \bar Q \rightarrow \pi\pi,~\eta\eta,~\eta\eta^\prime,~K\bar K$\\ \begin{eqnarray} M^{(0)SSC}_{Q\bar Q \rightarrow BC}& = & \sqrt{\frac{E_BE_C}{M_{f_0}}}~ \frac{8}{\pi^{1/4}}~\frac{R_{Q\bar Q}^{5/2}R_B^3}{(R_{Q\bar Q}^2+2R_B^2)^{5/2}}~ \exp\left\{ -\frac{1}{4}\left( \frac{R_{Q\bar Q}^2R_B^2}{R_{Q\bar Q}^2+2R_B^2}\right)~K^2\right\} \\ & & \left[ K^{1/2}-\frac{R_B^2(R_{Q\bar Q}^2+R_B^2)}{3(R_{Q\bar Q}^2+2R_B^2)}~ K^{5/2}\right]\nonumber \end{eqnarray} 2) $Q\bar Q \rightarrow \rho\rho,~\omega\omega$\\ \begin{eqnarray} M^{(l_{BC})SSC}_{Q\bar Q \rightarrow BC}& = & \sqrt{\frac{E_BE_C}{M_{f_0}}}~ \frac{8}{\sqrt{3}\pi^{1/4}}~\frac{R_{Q\bar Q}^{5/2}R_B^3}{(R_{Q\bar Q}^2+2R_B^2)^{5/2}}~ \exp\left\{ -\frac{1}{4}\left(\frac{R_{Q\bar Q}^2R_B^2}{R_{Q\bar Q}^2+2R_B^2}\right)~K^2\right\} \\ & & \cdot \left\{ \begin{array}{cc} K^{1/2}-\frac{R_B^2~(R_{Q\bar Q}^2+R_B^2)}{3~(R_{Q\bar Q}^2+2R_B^2)}~K^{5/2} & \mbox{for} \qquad l_{BC}=0 \\ & \\ -~\sqrt{\frac{8}{9}}~\cdot~ \frac{R_B^2(R_{Q\bar Q}^2+R_B^2)}{R_{Q\bar Q}^2+2R_B^2}~K^{5/2} & \mbox{for} \qquad l_{BC}=2 \end{array} \right. \end{eqnarray} 3) $Q\bar Q \rightarrow \sigma\sigma$ \\ \begin{eqnarray} M^{(0)SSC}_{Q\bar Q \rightarrow \sigma\sigma}& = & \sqrt{\frac{E_\sigma E_\sigma}{M_{f_0}}}~ \frac{4}{9\pi^{1/4}}~\frac{R_{Q\bar Q}^{5/2}R_B^5}{(R_{Q\bar Q}^2+2R_B^2)^{7/2}}~ \exp\left\{ -\frac{1}{4}\left( \frac{R_{Q\bar Q}^2R_B^2}{R_{Q\bar Q}^2+2R_B^2}\right)~K^2\right\} \\ & &\left[60K^{1/2}+\frac{11R_{Q\bar Q}^4+4R_B^2(R_{Q\bar Q}^2+R_B^2)}{R_{Q\bar Q}^2+2R_B^2}~K^{5/2} -\frac{R_{Q\bar Q}^4R_B^2(R_{Q\bar Q}^2+R_B^2)}{(R_{Q\bar Q}^2+2R_B^2)^2}~K^{9/2}\right] \nonumber \end{eqnarray} 4) $Q\bar Q \rightarrow \pi^*(1300)\pi$\\ \begin{eqnarray} M^{(0)SSC}_{Q\bar Q \rightarrow \pi^*(1300)\pi}& = & \sqrt{\frac{E_{\pi}E_{\pi^*}}{M_{f_0}}}~ \frac{2^{3/2}}{3^{3/2}\pi^{1/4}}~\frac{R_{Q\bar Q}^{5/2}R_B^3}{(R_{Q\bar Q}^2+2R_B^2)^{7/2}}~\exp\left\{ -\frac{1}{4}\left( \frac{R_{Q\bar Q}^2R_B^2}{R_{Q\bar Q}^2+2R_B^2}\right)~K^2\right\} \nonumber\\ & & \left[ 6\cdot (3R_{Q\bar Q}^2-4R_B^2)~K^{1/2}-\frac{R_{Q\bar Q}^2R_B^2(13R_{Q\bar Q}^2 +6R_B^2)}{R_{Q\bar Q}^2+2R_B^2}~K^{5/2} \right. \\ & & \left.~+~ \frac{R_{Q\bar Q}^4R_B^4(R_{Q\bar Q}^2 +R_B^2)}{(R_{Q\bar Q}^2+2R_B^2)^2}~K^{9/2}\right] \nonumber \end{eqnarray} 5) $Q\bar Q \rightarrow a_1\pi$\\ \begin{eqnarray} M^{(1)SSC}_{Q\bar Q \rightarrow a_1\pi}& = & \sqrt{\frac{E_{a_1}E_\pi}{M_{f_0}}}~ \frac{2^4}{3\pi^{1/4}} \frac{R_{f_0}^{5/2}R_B^4}{(R_{Q\bar Q}^2+2R_B^2)^{5/2}} \exp\left\{ -\frac{1}{4}\left( \frac{R_{Q\bar Q}^2R_B^2}{R_{Q\bar Q}^2+2R_B^2}\right)~K^2\right\}K^{3/2} \end{eqnarray} \newpage \section{Transformation brackets} The relevant values for the transformation brackets $<nl,NL,g|n_1l_1,n_2l_2,g>$ used in Eq. (19) are: \\ 1) $g=0$ \begin{eqnarray} <00,10,0|01,01,0> & = & \frac{1}{\sqrt{2}} \\ <01,01,0|01,01,0> & = & 0 \\ <10,00,0|01,01,0> & = & -\frac{1}{\sqrt{2}} \end{eqnarray} 2) $g=1$ \begin{eqnarray} <01,01,1|01,01,1> & = & 1 \end{eqnarray} 3) $g=2$ \begin{eqnarray} <00,02,2|01,01,2> & = & \frac{1}{\sqrt{2}} \\ <01,01,2|01,01,2> & = & 0 \\ <02,00,2|01,01,2> & = & - \frac{1}{\sqrt{2}} \end{eqnarray} \newpage \section{Digluonium decay amplitudes} The spin-spatial-color matrix elements $M^{SSC}_{G_0 \rightarrow BC}$, defined in Eq. (22), for the decay of the scalar digluonium state $G_0$ (size parameter $R_{G_0}$) into the final state mesons B C (size parameter $R_B$) are: \\ 1) $G_0 \rightarrow \pi\pi,~\eta\eta,~\eta\eta^\prime,~K\bar K~( \equiv 2~PS)$ \begin{eqnarray} M^{(0)SSC}_{G_0 \rightarrow 2PS} & = & \sqrt{\frac{E_BE_C}{M_{f_0}}}~ \frac{32}{3}~\sqrt{2}~\pi^{7/4}~\frac{R_{G_0}^{3/2}~(R_B^2-2R_{G_0}^2)}{(R_B^2+2R_{G_0}^2)^{5/2}}~K^{1/2} \end{eqnarray} 2) $G_0 \rightarrow \rho\rho,~\omega\omega~( \equiv 2~VM)$ \begin{eqnarray} M^{(0)SSC}_{G_0 \rightarrow 2VM} & = & \frac{1}{\sqrt{3}}~ M^{(0)SSC}_{G_0 \rightarrow 2PS} \end{eqnarray} 3) $G_0 \rightarrow \sigma\sigma$ \begin{eqnarray} M^{(0)SSC}_{G_0 \rightarrow \sigma\sigma} & = & \sqrt{\frac{E_\sigma E_\sigma}{M_{f_0}}}~ \frac{64\sqrt{2}}{9}~\pi^{7/4}~ \frac{R_{G_0}^{7/2}~(2R_{G_0}^2-9R_B^2)}{(R_B^2+2R_{G_0}^2)^{7/2}}~K^{1/2} \end{eqnarray} 4) $G_0 \rightarrow \pi^*(1300)\pi$ \begin{eqnarray} M^{(0)SSC}_{G_0 \rightarrow \pi^*\pi} & = & \sqrt{\frac{E_\pi^* E_\pi}{M_{f_0}}}~ \frac{32}{3\sqrt{3}}~\pi^{7/4}~ \frac{R_{G_0}^{3/2}~R_B^2~(14R_{G_0}^2-3R_B^2)}{(R_B^2+2R_{G_0}^2)^{7/2}}~K^{1/2} \end{eqnarray} \end{appendix} \newpage \include{reference} \newpage \begin{figure} \caption{\label{glueb}Leading order decay mechanism of the $G_0 \rightarrow G_0 G_0$ transition. The dot indicates the three-gluon coupling.} \end{figure} \begin{figure} \caption{\label{quarkonia}Quark line diagram for the decay of the quarkonia component $Q\bar Q$ into the final state mesons BC occurring as a leading order decay mechanism in strong coupling QCD. The dot indicates the $^3P_0$ vertex with strength $\lambda$.} \end{figure} \begin{figure} \caption{\label{gluea}Transition $G_0 \rightarrow BC$ occurring as a next-to-leading order decay mechanism in strong coupling QCD. The dots indicate the quark-gluon coupling with strength $\gamma$.} \end{figure} \begin{figure} \caption{\label{loword} Dependence of the branching ratios $BR(f_0(1500) \rightarrow BC)$ on the mixing angle $\alpha$, defined by the $Q\bar Q$ configuration $|f_0(1500)>=cos\alpha |n\bar n>-sin\alpha |s\bar s>$, for the decay mechanism $Q\bar Q \rightarrow BC$ of Fig. 2. For the meson size parameters we take the values given in the main text. The arrows indicate the predictions for the mixing schemes of Refs. \protect\cite{ams,weina} in lowest order of the decay mechanism.} \end{figure} \begin{figure} \caption{\label{width}Dependence of the total width of the $f_0(1500)$ resonance on the mixing angle $\alpha$ defined by the $Q\bar Q$ configuration $|f_0(1500)>=cos\alpha |n\bar n>-sin\alpha |s\bar s>$. The shadowed band corresponds to the experimental total width of $\Gamma=111\pm12~MeV$ \protect\cite{amsler}.} \end{figure} \begin{figure} \caption{\label{gluemass} Dependence of the branching ratios $BR(G_0 \rightarrow BC)$ on the glueball mass for the decay mechanism $G_0 \rightarrow BC$ of Fig. 3. Exact flavor $SU(3)$ symmetry for the quark-gluon vertex is assumed. For the size parameters we use the values given in the main text. For mass averaging we take a glueball width of $\Gamma_{G_0}=120~MeV$. } \end{figure} \newpage \begin{table} \caption{ Flavor matrix elements for the two-body decay $Q\bar Q \rightarrow BC$ of the mixed $f_0(M)$ states as defined in Eq. (11). The quarkonia component of the $f_0$ states is defined as $|Q\bar Q>=a_{n\bar n} |n\bar n>+a_{s\bar s}|s\bar s>$. The quantities $\alpha_{PS}$ and $\beta_{PS}$ are related to the pseudoscalar mixing angle as defined by Eq. (13). For the vector mesons we use ideal mixing.} \begin{tabular}{lcl} {\Large B~C} & & {\Large $M_{Q\bar Q \rightarrow BC}^F$}\\ \hline $ \pi\pi,\rho\rho$ & & $\sqrt{3}~a_{n\bar n}$ \\ $ \sigma\sigma,\omega\omega$ & & $a_{n\bar n}$ \\ $ \pi^*(1300)\pi,a_1\pi$ & & $\sqrt{6}~a_{n\bar n}$ \\ $ K\bar K$ & & $ a_{n\bar n}+\sqrt{2}~a_{s\bar s}$ \\ $ \eta\eta$ & & $\alpha_{PS}^2~a_{n\bar n}~+~\sqrt{2}~ \beta_{PS}^2~a_{s\bar s}$ \\ $\eta\eta^\prime$ & & $2~\alpha_{PS}~\beta_{PS}~(a_{n\bar n} /\sqrt{2}-a_{s\bar s})$ \end{tabular} \label{tab1} \end{table} \newpage \begin{table} \caption{ Flavor matrix elements of Eq. (21) for the two-body decay $G_0 \rightarrow BC$ of the mixed $f_0(M)$ states with glueball component $a_{G_0} |G_0>$. The quark-gluon coupling is assumed to be flavor independent.} \begin{tabular}{lcl} {\Large B~C} & & {\Large $M_{G_0 \rightarrow BC}^F/a_{G_0}$}\\ \hline $ \pi\pi,\rho\rho$ & & $\sqrt{6}$ \\ $ \sigma\sigma,\omega\omega$ & & $\sqrt{2}$ \\ $ \pi^*(1300)\pi,a_1\pi$ & & $\sqrt{12}$ \\ $ K\bar K$ & & $ 2~\sqrt{2}$ \\ $ \eta\eta$ & & $ \sqrt{2}$ \\ $ \eta\eta{^\prime} $ & & 0 \end{tabular} \label{tab2} \end{table} \newpage \begin{table} \caption{ Squared decay amplitudes and interference terms as defined by Eq. (5) for the transition $f_0(1500) \rightarrow BC$. Mass averaging as in Eq. (6) is already included. For the interference term we have $l_{BC}=0$ as restricted by the dynamical selection rule of the $G_0 \rightarrow BC$ transition. All values are given in MeV.} \begin{tabular}{llll} {\Large BC}& {\Large $M_{Q\bar Q \rightarrow BC}^2$} & {\Large $ M_{G_0 \rightarrow BC}^2$} & {\Large $ M_{G_0 \rightarrow BC}^{(0)} M_{Q\bar Q \rightarrow BC}^{(0)}$} \\ \hline $\pi\pi$ & $52.2a_{n\bar n}^2$ & $224.4a_{G_0}^2$ & $-103.1~a_{n\bar n} a_{G_0}$ \\ $\eta\eta$ & $20.9 (0.63a_{n\bar n}+0.523a_{s\bar s})^2$& $ 865.8 b_{G_0}^2$ & $-131.1a_{G_0} (0.63a_{n\bar n}+0.523a_{s\bar s})$ \\ $\eta\eta^\prime$ & $10.8(a_{n\bar n}/\sqrt{2}-a_{s\bar s})^2$ & 0 & 0 \\ $ K\bar K$ &$ 10.7 (a_{n\bar n}+\sqrt{2}a_{s\bar s})^2$ & $828.0 a_{G_0}^2$ & $-86.8 a_{G_0}(a_{n\bar n}+\sqrt{2}a_{s\bar s})$ \\ $\rho\rho$ & $50.4 a_{n\bar n}^2$ & $15.0 a_{G_0}^2$ & $10.2 a_{n\bar n} a_{G_0}$ \\ $\omega\omega$ & $5.3 a_{n\bar n}^2$ & $ 2.2 a_{G_0}^2$& $2.0 a_{n\bar n} a_{G_0}$ \\ $a_1\pi$ & $94.2 a_{n\bar n}^2$ & 0 & 0 \\ $\sigma\sigma$ & $16.1 a_{n\bar n}^2$ & $264.6 a_{G_0}^2$ & $-63.1 a_{n\bar n} a_{G_0}$ \\ $\pi^*(1300)\pi$ & $7.2 a_{n\bar n}^2$ & $440.4 a_{G_0}^2$& $-55.2 a_{n\bar n} a_{G_0}$ \end{tabular} \label{tab3} \end{table} \newpage \begin{table} \caption{ Predictions of mixing scheme A \protect\cite{ams} for the partial decay widths $\Gamma(BC)$ of the physical $f_0$ states, B C are the final state mesons. The $4\pi$ decay channel refers to the contributions of the $\sigma\sigma$ and $\pi^\ast\pi$ intermediate states using $BR(\pi^\ast \rightarrow \sigma \pi)=0.1-1$. All values are given in MeV. The predicted ratios $BR(f_0(1370) \rightarrow K\bar K)/BR(f_0(1370) \rightarrow \eta\eta)=2.6$ and $BR(f_0(1710) \rightarrow \pi\pi)/BR(f_0(1710) \rightarrow K\bar K)=0.93$ have to be compared to the experimental values of $\approx 2.5$ \protect\cite{amsler} and $0.39\pm0.14$ \protect\cite{pdg}.} \begin{tabular}{lccccccccccc} &$\Gamma_{tot}$ & $\pi\pi$ & $K\bar K$ & $\eta\eta$& $\eta\eta^\prime$& $\rho\rho$ & $ \omega\omega$ & $a_1\pi$ & $\pi^*\pi$&$ \sigma\sigma$ & $4\pi$ \\ \hline && & \multicolumn{4}{c}{{\large $f_0(1500)$}} & & & & \\ Model &136 & 29.3& 3.9& 5.2 & 19 & 19 & 2 & 36.8 & 9.9 & 12.5& 13.5-22.4 \\ Exp.\cite{amsler} & $111\pm12$ & $32\pm8.4$ & $3.9\pm0.3$ & $5.2\pm1.5$ & $1.4\pm0.4$& & & & & & $68\pm10.7$ \\ \hline && & \multicolumn{4}{c}{{\large $f_0(1370)$}} & & & & \\ Model & 410& 65.4 & 29.1& 10.8 & 3 & 112.1 & 22.8 & 138.6& 12.7 & 16 & 17.3-28.7 \\ Exp.\cite{amsler} & $351\pm41$ & &\multicolumn{2}{c}{$K\bar K/\eta\eta \approx 2.5$} &&&&&& \\ \hline & & & \multicolumn{4}{c}{{\large $f_0(1710)$}}& & & & \\ Model & 179 & 23.4 & 25.1 & 22.6 & 20.3 & 21.3 & 4.3 & 28.2 & 20.9 & 12.7 & 14.8-33.6 \\ Exp.\cite{pdg} & $133\pm14$ &\multicolumn{3}{c}{$\pi\pi/K\bar K=0.39\pm0.14$}&&&&&& \end{tabular} \label{tab4} \end{table} \newpage \begin{table} \caption{ Predictions of mixing scheme B \protect\cite{weina} for the partial decay widths $\Gamma(BC)$ of the physical $f_0$ states. The $4\pi$ decay channel refers to the contributions of the $\sigma\sigma$ and $\pi^\ast\pi$ intermediate states using $BR(\pi^\ast \rightarrow \sigma \pi)=0.1-1$. All values are given in MeV. The predicted ratios $BR(f_0(1370) \rightarrow K\bar K)/BR(f_0(1370) \rightarrow \eta\eta)=1.45$ and $BR(f_0(1710) \rightarrow \pi\pi)/BR(f_0(1710) \rightarrow K\bar K)=0.60$ have to be compared to the experimental values of $\approx 2.5$ \protect\cite{amsler} and $0.39\pm0.14$ \protect\cite{pdg}.} \begin{tabular}{lccccccccccc} &$\Gamma_{tot}$ & $\pi\pi$ & $K\bar K$ & $\eta\eta$& $\eta\eta^\prime$& $\rho\rho$ & $ \omega\omega$ & $a_1\pi$ & $\pi^*\pi$& $ \sigma\sigma$ & $4\pi$ \\ \hline & & & & \multicolumn{4}{c}{{\large $f_0(1500)$}} & & & \\ Model &157& 34.7& 3.9& 5.1 & 31.8 & 15.8 & 1.6 & 31.8 & 16.2 & 17.4& 19-33.6 \\ Exp. \cite{amsler} & $111\pm12$ & $32\pm8.4$ & $3.9\pm0.3$ & $5.2\pm1.5$ & $1.4\pm0.4$& & & & & & $68\pm10.7$ \\ \hline & &&& \multicolumn{4}{c}{{\large $f_0(1370)$}} & & & \\ Model &370 & 38.7 & 31.4& 21.6 & 1.3 & 110.8 & 22.6 & 132.2& 8.2 & 2.8 & 3.6-11 \\ Exp. \cite{amsler} & $351\pm41$ & &\multicolumn{2}{c}{$K\bar K/\eta\eta \approx 2.5$}&&&&&& \\ \hline & & & &\multicolumn{4}{c}{{\large $f_0(1710)$}} & & & \\ Model &2000 & 253 & 381 & 442 & 0.4 & 63 & 12& 76& 536 & 263 &316-799 \\ Exp. \cite{pdg} & $133\pm14$ &\multicolumn{3}{c}{$\pi\pi/K\bar K=0.39\pm0.14$}&&&&&& \end{tabular} \label{tab5} \end{table} \newpage \begin{table} \caption{ Predictions for the partial decay widths $\Gamma(BC)$ of the physical $f_0$ states for the extracted mixing scheme. The $4\pi$ decay channel refers to the contributions of the $\sigma\sigma$ and $\pi^\ast\pi$ intermediate states using $BR(\pi^\ast \rightarrow \sigma \pi)=0.1-1$. All values are given in MeV. The predicted ratios $BR(f_0(1370) \rightarrow K\bar K) /BR(f_0(1370) \rightarrow \eta\eta)=2.7$ and $BR(f_0(1710) \rightarrow \pi\pi)/BR(f_0(1710) \rightarrow K\bar K)=0.57$ have to be compared to the experimental values of $\approx 2.5$ \protect\cite{amsler} and $0.39\pm0.14$ \protect\cite{pdg}.} \begin{tabular}{lccccccccccc} &$\Gamma_{tot}$ & $\pi\pi$ & $K\bar K$ & $\eta\eta$& $\eta\eta^\prime$& $\rho\rho$ & $ \omega\omega$ &$a_1\pi$& $\pi^*\pi$&$ \sigma\sigma$ & $4\pi$ \\ \hline & && & \multicolumn{4}{c}{{\large $f_0(1500)$}} & & & \\ Model & 101& 23& 2.8& 6.7 & 14.6 & 9.7 & 1 & 19.6 & 11.8 & 12.1& 13.3-23.9 \\ Exp. \cite{amsler} & $111\pm12$ & $32\pm8.4$ & $3.9\pm0.3$ & $5.2\pm1.5$ & $1.4\pm0.4$& & & & & & $68\pm10.7$ \\ \hline && &&\multicolumn{4}{c}{{\large $f_0(1370)$}} & & & \\ Model &479 & 78.2 & 27.4& 10.1 & 4.6 & 133.2 & 27.1&164.9 & 14.6 & 19 & 20.5-33.6 \\ Exp. \cite{amsler} & $351\pm41$ & &\multicolumn{2}{c}{$K\bar K/\eta\eta \approx 2.5$}&&&&&& \\ \hline & && & \multicolumn{4}{c}{{\large $f_0(1710)$}} & & & \\ Model & 146 & 15 & 26.3 & 24.5 & 27.7 & 9.7 & 1.9&13& 18 & 10.1 &11.9-28.1 \\ Exp. \cite{pdg} & $133\pm14$ &\multicolumn{3}{c}{$\pi\pi/K\bar K=0.39\pm0.14$}&&&&&& \end{tabular} \label{tab6} \end{table} \newpage \centerline{Fig. 1} \centerline{\epsfbox{leadgluon.eps}} \newpage \centerline{Fig. 2} \centerline{\epsfbox{quarkonia.eps}} \newpage \centerline{Fig. 3} \centerline{\epsfbox{glue.eps}} \newpage \centerline{Fig. 4} \centerline{\epsfbox{f01500.eps}} \newpage \centerline{Fig. 5} \centerline{\epsfbox{totwidth.eps}} \newpage \centerline{Fig. 6} \centerline{\epsfbox{masgluedep.eps}} \newpage \end{document}
\section{Introduction} The thermal radiation from a black hole was first predicted by Hawking \cite{Hawking}, which phenomenon became widely known as the Hawking radiation. This prediction is based on quantum field theory in curved space, which is thought of as an effective theory valid for low energy physics. However, when we consider the mechanism of the Hawking radiation, crucial role is played by wave packets which left the past null infinity with very high frequency. Such wave packets propagate through the collapsing body just before the horizon formed, and undergo a large redshift on the way out to the future null infinity. Here arises one question. Can it be justified to apply quantum field theory in curved space, an effective theory at low energy, to the phenomenon which involves the infinitely high frequency regime? There may exist some unknown physics which invalidates the application of the standard quantum field theory in curved space \cite{BD}. One of such possibilities is that the spacetime may reveal its discrete nature at such high frequencies. To take account of the effect of possible modification of theory in the high frequency regime, Unruh proposed a simple toy model by a sonic analog of a black hole \cite{Unruh,Unruh1}. In Unruh model, the dispersion relation of fields at high frequencies are modified so as to eliminate very short wavelength modes. In some sense, this modification is arranged to reflect the atomic structure of fluid which propagates sound waves. Usually, the group velocity of sound waves drops to values much less than the low frequency value when the wavelength becomes comparable to the atomic scale. In performing such modifications \cite{Unruh1}, one must assume the existence of a reference frame because the concept of high frequencies can never be a Lorentz invariant one. Namely, the Unruh's model manifestly breaks the Lorentz invariance. To our surprise, even with such a drastic change of theory, the spectrum observed at the future infinity was turned out to be kept unaltered \cite{Unruh1,JacCor,Corley}. Here, in this paper, we consider a generalization of this model. The Lorentz invariance is the very basic principle for both the special relativity and the general relativity. Hence, there are many efforts to examine the violation of the Lorentz invariance \cite{will}, and new ideas to make use of high energy astrophysical phenomena are also proposed recently \cite{phili}. However, we have not had any evidence suggesting this rather radical possibility yet. Therefore, one may think that it is not fruitful to study in detail such a toy model that violates the Lorentz invariance at the moment. But, we also have another motivation to study this toy model even if we could believe that the Lorentz invariance is an exact symmetry of the universe. In most of literature, the Hawking radiation was studied in the framework of non-interacting quantum fields in curved space. However, when we consider a realistic model, it will be necessary to consider fields with interaction terms \cite{Hoo}. The evolution of interacting fields in the background that is forming a black hole is a very interesting issue but to study it is very difficult. Hence, as a first step, it will be interesting to take partly into account the interaction between the quantum fields and the matter which is forming a black hole. Then, it will be natural to introduce a modified dispersion relation associated with the rest frame of the matter remaining around the black hole. In this sense, the Unruh's model does not require that the fundamental theory itself violate the Lorentz invariance. In order to introduce the modified dispersion relation we need to specify one special reference frame. In previous works \cite{Unruh1,JacCor,Corley,BMPS}, the rest frame of freely-falling observers was employed as the special reference frame. In this case, the thermal spectrum of the Hawking radiation was shown to be reproduced. However, it is still uncertain whether the same thing remains true even when we adopt another reference frame as the special reference frame. In this paper, we give a generalization of previous works \cite{Unruh1,JacCor,Corley,BMPS} by allowing a more general choice of the reference frame. In most parts of the present paper, we follow the strategy taken in the paper by Corley\cite{Corley}. (See also \cite{BMPS}.) In his paper, as modifications of the Unruh's original model, two types of models were investigated. One is that with subluminal dispersion relation and the other is that with superluminal dispersion relation. It was shown analytically that the thermal spectrum at the Hawking temperature is reproduced in both cases. However, in the superluminal case, the standard notion of the causal structure of black hole breaks down. Even if we consider the case that the background geometry is given by a Schwarzschild black hole, the wave packets corresponding to the radiated particles can be traced back to the singularity inside the horizon due to their superluminal nature. Hence, the singularity becomes naked, and we confront the problem of the boundary condition at the singularity. To avoid this difficulty, it is often required that the vacuum fluctuations be in the ground state just inside the horizon. However, it is not clear what is the correct boundary condition. As a topic related to the superluminal dispersion relation, it was also reported that the Hawking radiation is not necessarily reproduced in the models with an inner horizon \cite{CoJ}. In this paper, we wish to focus on the subluminal case leaving such a delicate issue related to the superluminal dispersion relation for the future problem. Even in the restriction to the subluminal case, it will be important to study various models to examine the universality of the Hawking radiation. In the present paper, we finally find that the resulting spectrum of created particle stays thermal one at the Hawking temperature as long as we mildly change the choice of the special reference frame. By a systematic use of the technique of the so-called matched asymptotic expansion, we also evaluate how small the leading correction to the thermal radiation is. On the other hand, for some extreme modification of the reference frame, in which case the analytic treatment is no longer valid, the spectrum is numerically shown to deviate from the thermal one significantly. This paper is organized as follows. First we introduce a generalization of the Unruh's model in Sec.2. In Sec.3 we review what quantities need to be evaluated in computing the spectrum of particle creation in our model. In Sec.4, we construct a solution of the field equation, and we evaluate the spectrum of created particles by using this solution. To determine the order of the leading correction to the thermal spectrum, we employ the method of asymptotic matching in Sec.5. In Sec.5, we also demonstrate some results of numerical calculation to verify our analytic results. In addition, we display the results for some extreme cases which are out of the range of validity of our analytic treatment. Section 6 is devoted to conclusion. Furthermore, appendix C is added to discuss the effect of scattering due to the modified dispersion relation. Although we do not think that this effect is directly related to the issue of Hawking radiation, it can in principle change the observed spectrum of Hawking radiation drastically if it accumulates throughout the long way to a distant observer. In the present paper, we use units with $ \hbar = c = G = 1$. \section{Model} In this section, we explain how we generalize the model that was investigated in the earlier works \cite{Unruh1,JacCor,Corley}. Following these references, we consider a massless scalar field propagating in a 2-dimensional spacetime given by \begin{equation} ds^2=-d\tilde t^2+(d\tilde x-\tilde v(\tilde x)d\tilde t)^2, \label{orimet} \end{equation} where $\tilde v(x)$ is a function which goes to a constant at $\tilde x\to\infty$ and satisfies $\tilde v(\tilde x)\geq -1$ for $\tilde x\geq 0$. The equality holds at $\tilde x=0$. Since the line element $d\tilde x=0$ is null at $\tilde v=-1$, we find that the event horizon is located at $\tilde v=-1$. Furthermore, by the coordinate transformation given by $d\tilde t=d\tilde t'+\tilde v/(1-\tilde v^2)d\tilde x$, the above metric can be rewritten as \begin{equation} ds^2=-(1-\tilde v^2)d\tilde t'{}^2+{1\over 1-\tilde v^2} d\tilde x^2. \label{met2} \end{equation} If we set $\tilde v(\tilde x) =-\sqrt{2M/(\tilde x+2M)}$, this metric represents a 2-dimensional counterpart of a Schwarzschild spacetime with the event horizon at $\tilde x=0$. In this coordinate system, the unit vector perpendicular to the $\tilde t=$constant hypersurfaces is given by $\tilde u_{\alpha}:=\partial_{\alpha} \tilde t$, and the differentiation in this direction is given by $\tilde u^{\alpha}\partial_{\alpha}= (\partial_{\tilde t}+\tilde v\partial_{\tilde x})$. We denote the unit outward pointing vector normal to $\tilde u_{\alpha}$ by $\tilde s^{\alpha}$. In order to examine the effect on the spectrum of the Hawking radiation due to a modification of theory in the high frequency regime, they investigated a system defined by the modified action of a scalar field, \begin{equation} S={1\over 2}\int d^2\tilde x\sqrt{-g}\, g^{\alpha\beta} {\cal D}_{\alpha}\phi^{*}{\cal D}_{\beta}\phi, \label{act1} \end{equation} where the differential operator ${\cal D}$ is defined by $\tilde u^{\alpha} {\cal D}_{\alpha} = \tilde u^{\alpha} \partial_{\alpha}, \tilde s^{\alpha}{\cal D}_{\alpha} = \hat F(\tilde s^{\alpha}\partial_{\alpha})$. If we set $\hat F(z)=z$, the action (\ref{act1}) reduces to the standard form. Since we are interested in the effect caused by the change in the high frequency regime, we assume that $\hat F(z)$ differs from $z$ only for large $z$. In the above model, the dispersion relation for the scalar field manifestly breaks the Lorentz invariance, and there is a special reference frame specified by $\tilde u$. One can easily show that this reference frame is associated with a set of freely-falling observers. As noted in Introduction, it was shown that the spectrum of the Hawking radiation is reproduced in this model. Here we consider a further generalization of this model allowing to adopt other reference frames as the special reference frame. However, because of technical difficulties, we restrict our consideration to stationary reference frames. As we are working on a 2-dimensional model, the reference frame is perfectly specified by choosing one time-like unit vector, which we denote by $u$. Since $\partial_{\tilde t}$ in the original coordinate system $(\tilde t, \tilde x)$ is a time-like Killing vector, the condition for the reference frame to be stationary becomes $\lee_{\partial_{\tilde t}} u^{\alpha}=0$, where $\lee_{\partial_{\tilde t}}$ is the Lee derivative in the direction of $\partial_{\tilde t}$. This condition can be simply written as ${\partial u^{\tilde \alpha}/\partial{\tilde t}}=0$, where we used indices associated with $\tilde{}$ to represent the components in the $(\tilde t,\tilde x)-$coordinate. Furthermore, the covariant components $u_{\tilde \alpha}(\tilde x) =g_{\tilde \alpha \tilde \beta}(\tilde x) u^{\tilde \beta}(\tilde x)$ is also independent of $\tilde t$. Thus, if we introduce a new time coordinate $t$ by \begin{equation} dt=u_{\tilde 0}^{-1}(\tilde x)\left( u_{\tilde 0}(\tilde x) d\tilde t+ u_{\tilde 1}(\tilde x) d\tilde x\right) =d\tilde t -\gamma (\tilde x)d\tilde x, \label{tconst} \end{equation} the $t-$constant hypersurfaces become manifestly perpendicular to $u_{\alpha}$. Here we introduced $\gamma(\tilde x):=-u_{\tilde 1}(\tilde x)/u_{\tilde 0}(\tilde x)$. Further, it is convenient to choose a new spatial coordinate $x$ so that $\partial_t$ coincide with the Killing vector Since \begin{equation} \partial_{t} \partial_{\tilde t}+{\partial x\over \partial \tilde t} \partial_{\tilde x}, \end{equation} $x$ should be chosen as a function which depends only on $\tilde x$. Hence, we set \begin{equation} x :=\int_0^{\tilde x} \zeta(\tilde x') d\tilde x'. \label{newx} \end{equation} By using such a new coordinate $(t,x)$ with \begin{equation} \zeta(\tilde x')=(1-\tilde v \gamma)^2-\gamma^2, \end{equation} the metric (\ref{orimet}) can be written in the form\footnote{ By considering sonic analog of the Hawking radiation, the use of this type of conformal metric was discussed \cite{Visser}.}, \begin{equation} ds^2= \Omega^2(x)\left[-dt^2+(dx-v(x) dt)^2\right], \label{newmet} \end{equation} where we set \begin{eqnarray} \Omega^2(x)&:=&{1\over (1-\tilde v \gamma)^2-\gamma^2}, \label{Omegadef} \\ v(x)&:=&\gamma+\tilde v(1-\tilde v \gamma). \label{newv} \end{eqnarray} Here we mention the constraint on $\Omega$. If we explicitly write down the condition that $u$ is a time-like unit vector, i.e., $u_\alpha u^\alpha=-1$, we find that $\Omega^2=u_{\tilde 0}^2$ holds. Hence $\Omega^2 >0$ is guaranteed as long as $u$ is time-like. Also, directly from the metric (\ref{newmet}), we can easily verify $\Omega^2>0$ if and only if the $t-$constant hypersurfaces are space-like. Hence, it will be appropriate to assume $\Omega^2>0$. Then, we find that $\Omega^2$ has a finite minimum value when $|\tilde v|<1$. The minimum value is $1-\tilde v^2$ which is realized when $v=0$. Next we write down the explicit form of $u$ and $s$. By using the facts that $u$ is perpendicular to the $t-$constant hypersurfaces and that it is an unit vector, we can show that the differentiation in the direction of $u$ is given by \begin{equation} u^{\alpha}\partial_{\alpha} ={1\over{\Omega}}\,(\partial_t+v\partial_x). \label{equ} \end{equation} Similarly, we can show that the differentiation in the direction of $s$, which is an unit vector perpendicular to $u$, is given by \begin{equation} s^{\alpha}\partial_{\alpha}={1\over{\Omega}}\,{\partial_x}. \end{equation} By using $u$ and $s$, the metric can be represented as $g^{\alpha\beta}=-u^{\alpha}u^{\beta}+s^{\alpha}s^{\beta}$, and the determinant of $g_{\alpha\beta}$ becomes $g=- \Omega ^4$. Now, we find that to generalize the choice of the special reference frame introduced to set the modified dispersion relation is equivalent to generalize the metric form given in (\ref{orimet}) to the one given in (\ref{newmet}) replacing $\tilde u$ and $\tilde s$ with $u$ and $s$ in the defining equations of the differential operator ${\cal D}$. As a result, the action (\ref{act1}) becomes \begin{equation} S={1\over 2}\int dtdx \left[ |(\partial_t+v\partial_x)\phi|^2- {{\Omega}(x)}^2 \left|{\hat F} \left({1\over{\Omega}}\partial_x \right) \phi \right|^{2} \right]. \label{action2} \end{equation} If we set $\Omega^2\equiv 1$, the models are reduced to the original one discussed in Ref.~\cite{Corley}. Here, we should mention one important relation for the later use. The temperature of the Hawking spectrum is determined by the surface gravity $\kappa$ defined by $\kappa:=d\tilde v/d\tilde x|_{\tilde x=0}$. The surface gravity $\kappa$ can also be represented as \cite{JacKan} \begin{equation} \kappa=\left.{dv\over dx}\right\vert_{x=0}. \label{temperature} \end{equation} In order to verify this relation, we evaluate $dv/dx$ by using Eqs.~(\ref{newx}) and (\ref{newv}) to obtain \begin{equation} {dv\over dx}={1\over {(1-\tilde v \gamma)^{2}-\gamma^{2}}} (\gamma_{,\tilde x}+\tilde v_{,\tilde x} (1-\tilde v\gamma)-\tilde v\tilde v_{,\tilde x} \gamma -\tilde v^{2}\gamma_{,\tilde x}). \label{surface} \end{equation} From Eq.~(\ref{newv}), we also find $v=-1$ when $x=0$. Hence, substituting $v=-1$ into (\ref{surface}), we obtain Eq.~(\ref{temperature}). Then, let us derive the field equation by taking the variation of the action (\ref{action2}). Assuming that ${\hat F}(z)$ is an odd function of $z$, we obtain \begin{equation} ({\partial}_{t}+{\partial}_{x}v)({\partial}_{t}+v{\partial}_{x}){\phi} ={\Omega}{\hat F}\!\left({1\over{\Omega}}{\partial}_{x} \right) {\Omega} \,{\hat F}\!\left({1\over{\Omega}}{\partial}_{x} \right){\phi}. \label{field equation} \end{equation} To proceed further calculations, we need to assume a specific dispersion relation. Following Ref.~\cite{Corley}, we adopt here \begin{equation} \hat F(z)=z+{1\over 2k_0^2} z^3, \label{dispersion} \end{equation} where $k_0$ is a constant. Since the model should be arranged to differ from the ordinary one only in the high frequency regime, the critical wave number $k_0$ is supposed to be sufficiently large. With this choice of $\hat F$, neglecting the terms that are inversely proportional to the fourth power of $k_0$, the field equation becomes \begin{equation} ({\partial}_{t}+{\partial}_{x}v)({\partial}_{t}+v{\partial}_{x}){\phi} =\left[\partial_x^2+{1\over 2 k_0^2} \left( \partial_x^2 {1\over \Omega} \partial_x {1\over \Omega} \partial_x \right)+ {1\over 2 k_0^2} \left(\partial_x{1\over \Omega} \partial_x {1\over \Omega} \partial_x^2\right)\right] \phi. \label{feq} \end{equation} Before closing this section, we briefly discuss the meaning of the functions $\Omega$ and $v$. From (\ref{equ}), it is easy to understand that $v$ is the coordinate velocity of the integration curves of $u$. To understand the meaning of $\Omega$, we further calculate the covariant acceleration of the integral curves of $u$, $|u^{\alpha}_{~;\beta} u^{\beta}|$, where semicolon represents the covariant differentiation. After a straightforward calculation, we see that the covariant acceleration is given by $|\partial_x \Omega^{-1}|$. Hence, we find that the derivative of $\Omega^{-1}$ gives the acceleration of the reference frame. \section{particle creation rate} In this section, we briefly review how to evaluate the spectrum of Hawking radiation. We clarify what quantities need to be calculated for this purpose. (For the details, see Ref.~\cite{JacCor}.) To evaluate the spectrum of Hawking radiation, we need to solve the field equation (\ref{feq}) with an appropriate boundary condition. However, owing to the time translation invariance with the Killing vector $\partial_t$, we do not have to solve the partial differential equation (\ref{feq}) directly. By setting $ {\phi}(t,x)=e^{-i{\omega}t}{\psi}(x)$, Eq.~(\ref{feq}) reduces to an ordinary differential equation (ODE), \begin{equation} \left[(-i\omega+{\partial}_{x} v)(-i\omega +v{\partial}_{x}) -\left\{{{\partial}_x}^{2}+{1\over {2 {k_0}^2}} {\left({{\partial}_{x}}^{2}{1\over{\Omega}} {\partial}_{x}{1\over{\Omega}} {\partial}_{x}\right)}+ {1\over{2 {k_0}^2}}{\left({\partial}_{x}{1\over{\Omega}}{\partial}_{x} {1\over{\Omega}}{{\partial}_{x}}^2\right)}\right\}\right] \psi =0. \label{ode} \end{equation} We could not make use of this simplification if we relax the restriction that the reference frame should be stationary. Here we note that the norm of $\partial_t$ is given by $|\partial_t|=\sqrt{1-v^2}\Omega= \sqrt{1-\tilde v^2}$. Therefore, the frequency $\omega^{(stat)}$ for the static observers who stay at a constant $x$(or $\tilde x$) is related to $\omega$ by \begin{equation} \omega^{(stat)}=\omega/\sqrt{1-\tilde v^2}. \label{omegas} \end{equation} Hence, as long as $\tilde v_{\infty}:=\tilde v(x\to\infty)$ is not equal to zero, $\omega^{(stat)}$ differs from $\omega$. By looking at the metric (\ref{met2}) in the static chart, we find that this frequency shift is merely caused by the familiar effect due to the gravitational redshift. Therefore, even if we consider models with $\tilde v_{\infty}\ne 0$, $\omega$ might be identified with the frequency observed at the hypothetical infinity where the gravitational potential is set to be zero. However, the situation is more transparent if we can set $\tilde v_{\infty}=0$ like the 2-dimensional black hole case. In this case, we can identify $\omega$ with $\omega^{(stat)}$ without any ambiguity. As mentioned in Ref. \cite{JacCor}, there is a difficulty in evaluating the spectrum of radiation in the case of $v_{\infty}:=v(x\to\infty)=0$. In previous models, $v_{\infty}=0$ directly implies $\tilde v_{\infty}=0$. Therefore, we could not apply the result to the example of a 2-dimensional black hole spacetime directly\footnote{ In the model proposed in Ref.~\cite{lattice}, the case with $v_{\infty}=0$ can be dealt with.}. On this point, in our extended model, the cases with $\tilde v_{\infty}=0$ can be dealt with since $\tilde v_{\infty}=0$ does not mean $v_{\infty}\ne 0$. Let us return to the problem to solve Eq.~(\ref{ode}). From the above ODE, the asymptotic solution at $x\to\infty$ is easily obtained by assuming the plane wave solution like \begin{equation} \psi(x)\propto e^{ikx}. \end{equation} Substituting this form into Eq.~(\ref{ode}), we obtain the dispersion relation \begin{equation} (\omega - v_{\infty}k)^{2} = k^{2} - {k^{4} \over k_0^{2} \Omega_{\infty}^2}, \label{disp} \end{equation} where $\Omega_{\infty}$ is the asymptotic constant value of $\Omega$. The quantity on the left hand side \begin{equation} \omega':=\omega-v k, \end{equation} is related to the frequency measured by the observers in the special reference frame. In fact, this frequency divided by $\Omega$ is the factor that appears when we act the operator $iu^{\alpha}\partial_\alpha$ on a wave function $e^{-i(\omega t-k x)}$. As shown in Ref.~\cite{JacCor}, two of the 4 solutions of Eq.~(\ref{disp}) have large absolute values, which we denote by $k_{\pm}$, and the other two have small absolute values, which we denote $k_{\pm s}$. For each pair, one is positive and the other is negative. The subscript $\pm$ represents the signature of the solution. Then, the general solution of Eq.~(\ref{disp}) at $x\to\infty$ is given by a superposition of these plane wave solutions as \begin{equation} \psi(x)=\sum_{l=\pm,\pm s} c_{l}(\omega)e^{ik_l(\omega)x}. \end{equation} By such a local analysis, however, the coefficients $c_{\ell}(\omega)$ are not determined. To determine them, we need to find the solution that satisfies the boundary condition corresponding to no ingoing waves plunging into the event horizon. This condition is slightly different from the condition that the solution of ODE~(\ref{ode}) vanishes inside the horizon. The latter condition is stronger than the former one because the latter one also prohibits the pure outgoing wave from the event horizon which may exist in the present model with the modified dispersion relation. The former condition is the sufficient condition to determine the wave function uniquely, while the existence of the solution that satisfies the latter condition is not guaranteed in general. However, once we find such a solution that satisfies the latter stronger condition, it is the solution that satisfies the required boundary condition. In the succeeding sections, we solve ODE~(\ref{ode}) requiring the latter condition. {}Finally, we present the formula to evaluate the expectation value of the number of emitted particles naturally defined at $x\to\infty$. In spite of our generalization of models, the same derivation of the formula that is given in Ref.~\cite{JacCor} is still valid. The same arguments follow without any change, but one possible subtlety exists on the point whether the expression of the conserved inner product is unaltered or not. Therefore, we briefly explain this point. The defining expression for the conserved inner product given in Ref.~\cite{JacCor} is \begin{equation} ({\phi_1},{\phi_2})=i{\int}dx[{\phi_1}^{*} ({\partial}_{t}+v{\partial}_{x}) {\phi_2}-{\phi_2}({\partial}_{t}+v{\partial}_{x}){\phi_1}^{*}], \label{inner} \end{equation} where the integration is taken over a $t$-constant hypersurface. Here both ${\phi_1}$ and ${\phi}_2$ are supposed to be solutions of the field equation. The constancy of this inner product is related to the invariance of the function \begin{equation} {\cal L}(\phi_1,\phi_2):=(\partial_t +\partial_x)\phi_1 \cdot (\partial_t +\partial_x)\bar\phi_2 -\Omega^2 \left[\hat F\left({1\over \Omega}\partial_x \right)\phi_1\right] \cdot \left[\hat F\left({1\over \Omega}\partial_x \right)\bar\phi_2\right], \end{equation} under the global phase transformation \begin{equation} \phi\to e^{i\lambda}\phi. \end{equation} Taking the differentiation of ${\cal L}(e^{i\lambda}\phi_1,e^{i\lambda}\phi_2)$ with respect to $\lambda$, we have \begin{eqnarray} 0={d{\cal L}(e^{i\lambda}\phi_1,e^{i\lambda}\phi_2) \over d\lambda}&=&{\partial {\cal L}\over \partial\phi_1}i\phi+ {\partial {\cal L}\over \partial\dot\phi_1}{d(i\phi_1)\over dt} +{\partial {\cal L}\over \partial\bar\phi_2}(-i\bar\phi_2)+ {\partial {\cal L}\over \partial\dot{\bar\phi_2}}{d(-i\bar\phi_2)\over dt} \cr &=&i{d\over dt}\left({\partial {\cal L}\over \partial\dot\phi_1}\phi_1- {\partial {\cal L}\over \partial\dot{\bar \phi_2}} {\bar\phi_2} \right), \label{constip} \end{eqnarray} where we used the field equation in the last equality. Eq.~(\ref{constip}) proves the constancy of the inner product (\ref{inner}). Of course, the constancy of the inner product (\ref{inner}) can also be verified by directly calculating its $t$-derivative using the field equation. Now that we verified that the extension of model does not alter the expression of the inner product, we just quote the formula from Ref.~\cite{JacCor}. For a wave packet which is peaked around a frequency $\omega$, the expectation value of the number of created particles is given by \begin{equation} N(\omega)={|\omega^{'}(k_-) v_g(k_-) c_-^{2}(\omega)| \over{|\omega^{'}(k_{+s}) v_g(k_{+s}) c_{+s}^{2}(\omega)|} }. \label{number} \end{equation} where $ v_g(k):=\partial\omega(k)/\partial k$ is the group velocity measured by a static observer. \section{solving field equation} In this section, to determine the coefficients $c_l$, we solve the field equation (\ref{ode}) by using several approximations. In the region close to the horizon, we use the method of Fourier transform. The solution is found to be uniquely determined by imposing the boundary condition discussed in the preceding section. On the other hand, in the region sufficiently far from the horizon, we construct four independent solutions which become $e^{ik_{\ell}(\omega) x}$ at $x\to \infty$. We use the WKB approximation for the two short-wavelength modes and we use the simple $1/k_0^2$ expansion for the other two long-wavelength modes. Later, we find that these two different regions of validity have an overlapping interval as long as $k_0$ is taken to be sufficiently large. Hence, the requirement that the solutions obtained in both regions match in this overlapping interval determines the coefficients $c_{\ell}(\omega)$. Basically, our computation is an extension of that given by Corley~\cite{Corley}. Here we take into account the generalization of models discussed in Sec.~2. Furthermore, to evaluate the order of the leading deviation from the thermal spectrum, we shall take into account some higher order terms. At the same time, we also carefully keep counting the order of errors contained in our estimation. For brevity, we concentrate on the most interesting case in which $\omega$ and $\kappa$ are same order. \subsection{the case close to the horizon} Now we want to find a solution satisfying the boundary condition that the wave function rapidly decrease in the horizon. Therefore, we restrict our consideration to the region close to the horizon, $|x|<x_1$, by choosing a sufficiently small $x_1$. We introduce a parameter, \begin{equation} \delta:=\max_{|x|<x_1, i=0, 1, 2} \tilde\delta_i(x), \label{ddelta} \end{equation} with \begin{equation} \tilde\delta_0:=\kappa x, \quad \tilde\delta_1:={\kappa_1^2 x\over \kappa}, \quad \tilde\delta_2:={\Omega_0\over \Omega_1} x. \end{equation} Since we wish to think of $\delta$ as a small parameter for the perturbative expansion, we require $\delta\ll 1$. Then, we find that $x_1$ must be chosen to satisfy \begin{equation} x_1\ll \mbox{min}\left(1/\kappa, \kappa/\kappa_1^2, \Omega_1/\Omega_0\right). \label{x_1} \end{equation} We assume that $v(x)$ and $1/\Omega(x)$ can be expanded around the horizon like \begin{eqnarray} v(x) & = & -1+\kappa x+{1\over 2} \kappa_1^2 x^2+O(\delta^3), \cr \left({1\over \Omega}\right) & \approx & {1\over \Omega_0} +{1\over \Omega_1} x +{1\over \Omega_0}O(\delta^2). \label{vomegat} \end{eqnarray} Substituting these expansions into the field equation (\ref{ode}), we classify the terms into five parts according to the number of differentiations acting on $\psi$. Then, keeping the leading order correction terms with respect to $\delta$ in each part, we obtain the field equation valid in the region close to the horizon as \begin{eqnarray} L[\psi(x)]: = && {1\over k_0^2} \left[{1\over \Omega_0^2} +{2\over \Omega_0 \Omega_1}x\right]\psi^{(4)} +{4\over k_0^2}{1\over \Omega_0 \Omega_1} \psi''' +(2\kappa x -\kappa^{2} x^{2}+\kappa_1^2 x^2) \psi'' \nonumber\\ && + 2\left[ -(i\omega - \kappa)+(\kappa (i \omega - \kappa ) + \kappa_1^2)x \right] \psi' - i \omega (i \omega - \kappa - \kappa_1^2 x) \psi = 0. \label{horizon feq} \end{eqnarray} In Corley's paper \cite{Corley}, the terms of $O(\delta^{1})$ were neglected, while we keep them in the present paper. Here, we introduce the momentum-space representation of the wave function, $\psi(s)$, by \begin{equation} \psi(x)=\int_{C}{\, }ds{\, }e^{sx} \hat{\psi}(s). \label{int} \end{equation} Substituting this expression into Eq.~(\ref{horizon feq}), we perform the integration by parts like \begin{eqnarray*} \int_C ds\, x e^{sx}\hat\psi(s)=\int_C ds\, \left({\partial\over\partial s} e^{sx}\right) \hat\psi(s)= - \int_C ds\, e^{sx} \left({\partial\over\partial s} \hat\psi(s)\right) +[e^{sx} \hat\psi(s)]_{s_i}^{s_f}, \end{eqnarray*} where $s_i$ and $s_f$ are the start and the end points of integration, respectively. Note that there appear surface terms like the last term on the right hand side in the above example. Then, we find that the field equation in the momentum space is given by \begin{equation} L[\psi(x)]:=\int_{C}{\, }ds{\, } {\hat L}[\hat \psi(s)] + \mbox{(boundary terms)} = 0, \label{fourier} \end{equation} where \begin{eqnarray} {\hat L}[\hat \psi(s)]:=&& {1\over k_0^2\Omega_0^2}\left(1- 2{\Omega_0\over \Omega_1}\left(\partial_s-{2\over s}\right)\right) s^4\hat\psi -\left(2\kappa \partial_s +(\kappa^2+\kappa_1^2)\partial_s^2 \right) s^2\hat\psi \nonumber\\ && +2\left[ -(i\omega - \kappa)-(\kappa(i\omega-\kappa)+ \kappa_1^2)\partial_s \right]s \hat\psi -i\omega(i\omega-\kappa+\kappa_1^2 \partial_s)\hat\psi. \label{bt} \end{eqnarray} If we are allowed to neglect the boundary terms, we can construct a solution of Eq.~(\ref{horizon feq}) by using Eq.~(\ref{int}) from a solution $\hat \psi(s)$ which satisfies \begin{equation} {\hat L}[\hat \psi(s)]=0, \label{bt1} \end{equation} In the following, we solve Eq.~(\ref{bt1}) without taking any care about whether the boundary terms can be neglected or not. After we find a solution, we verify that the corresponding boundary terms are sufficiently small. To solve Eq.~(\ref{bt1}), it is convenient to introduce a new variable \begin{equation} W:={\partial \ln(s^2\hat\psi)\over \partial \ln s}. \label{W1} \end{equation} Taking $\delta$ as a small parameter, we expand $W$ as \begin{equation} W=W^{(0)}+W^{(1)}+O(\delta^2). \label{Wexpansion} \end{equation} Then, $W^{(0)}$ is found to be given by \begin{equation} W^{(0)} = \left[{\tilde\epsilon^2\over 2}(sx)^3 +\left(1-i {\omega\over\kappa}\right)\right], \label{s1} \end{equation} and $W^{(1)}$ is given by \begin{equation} W^{(1)} = {\cal R}(W^{(0)}), \end{equation} with \begin{eqnarray} {\cal R}(W) & := & \tilde\delta_0 \left[ -{1\over 2}{\partial W\over \partial \ln s} -{1\over 2} W^2+{1\over 2}W + \left(1-{i\omega\over \kappa}\right) \left(W+{i\omega\over 2\kappa}-1\right) \right](sx)^{-1} -\tilde\delta_2 \tilde\epsilon^{2} W (sx)^2 \cr && -\tilde\delta_1 \left( -{1\over 2}{\partial W\over \partial \ln s}-{1\over 2} W^2 +{3\over 2}W -1 \right)(sx)^{-1}. \end{eqnarray} Here we defined \begin{equation} \tilde\epsilon:=\sqrt{1\over k_0^2\Omega_0^2\kappa x^3}. \label{ep2} \end{equation} Although we later consider the situation in which $\tilde\epsilon$ is small, $\tilde\epsilon$ is not small at all in the region close to the horizon. Substituting the explicit expression for $W^{(0)}$ into ${\cal R}(W)$, we obtain \begin{eqnarray} W^{(1)} &=& {\tilde\epsilon^4 (-\tilde\delta_0+\tilde \delta_1-4\tilde\delta_2) \over 8}(sx)^{5} -{\tilde\epsilon^2 \over 2}\left(\tilde\delta_0 +(\tilde\delta_1-2\tilde\delta_2) \left({i \omega \over \kappa}-1 \right) \right)(sx)^{2} \cr && + {i \omega \over {2 \kappa}} \tilde\delta_1 \left( {i \omega \over \kappa} + 1 \right)(sx)^{-1}. \end{eqnarray} Under the condition that the boundary terms in Eq.~(\ref{fourier}) can be neglected, we obtain the solution $\psi(x)$ which is valid up to $O(\delta^1)$ as \begin{equation} \psi(x) \approx \int_{C} ds\, e^{xf(s)}, \label{int2} \end{equation} where, \begin{equation} xf(s) = x f_0(s) + x f_1(s), \label{xfs} \end{equation} is given by \begin{eqnarray} x f_0(s) &:=& x s + \int {ds\over s} W^{(0)} -2\ln s \cr &=& xs +{ \tilde\epsilon^2 (xs)^{3} \over 6} +\left(-1-{iw\over \kappa}\right)\ln s, \label{xf0} \\ x f_1(s)&:=& \int {ds\over s} W^{(1)} \cr &=& {\tilde\epsilon^4 (-\tilde\delta_0 +\tilde\delta_1-4\tilde\delta_2) \over 40} (sx)^{5} -{\tilde\epsilon^2 \over 4} \left(\tilde\delta_0+(\tilde\delta_1-2 \tilde\delta_2) \left({i \omega \over \kappa}-1 \right)\right)(sx)^{2} \cr && - {i \omega \over {2 \kappa}} \tilde\delta_1 \left( {i \omega \over \kappa} + 1 \right)(sx)^{-1}. \label{xf1} \end{eqnarray} The boundary condition of the wave function requires that it exponentially decreases inside the horizon. In order to construct the wave function that satisfies this boundary condition, we must choose the contour of integration appropriately. We propose to adopt the contour $C$ given in Fig.~1. This contour does not go to infinity but have end points (denoted by open circles in Fig.~1) at which the absolute value of $s$ is sufficiently large but does not exceed the limit given in (\ref{region1}). Hence, as shown in Appendix \ref{appendixA}, the higher order correction does not dominate along this contour. Furthermore, as is also explained in Appendix \ref{appendixA}, with this choice of end points, the boundary terms in Eq.~(\ref{fourier}) are exponentially small, and can be neglected. \begin{figure}[tbh] \vspace*{0cm} \centerline{\epsfxsize5cm \epsfbox{himeron1.eps}} \caption{The contour for integration is chosen for the correction terms not to dominate the leading terms. In the shaded region, the higher order terms with respect to $\delta$ becomes larger than the leading terms. Hence, the contour is shortened so as not to violate the validity of approximation. } \label{fig1} \end{figure} Next we show that $\psi(x)$ given by (\ref{int2}) is actually the solution that satisfies the required boundary condition. In order to evaluate the integration (\ref{int2}) analytically, we use the method of steepest descents. For this method to be valid, $1/|x s_0|\ll 1$ is required, where $s_0$ is the value of $s$ at the saddle point that dominates the integral. Then, this requirement can be rewritten as \begin{equation} |x|> x_2, \end{equation} where we must choose $x_2$ to satisfy $(1/|x s_0|)_{x=x_2}\ll 1$. We shall see later that $1/|x s_0|$ is of $O(|\tilde\epsilon|)$. Hence, for the first time at this moment, we further restrict our consideration to the region in which $\tilde\epsilon$ is also small. In order that the condition $|x|>x_2$ to be compatible with $|x|<x_1$, \begin{equation} \left({\kappa\over k_0\Omega_0}\right)^{2/3} \ll \mbox{min}(1, \kappa^2/\kappa_1^2, \kappa\Omega_1/\Omega_0), \label{matchcondi} \end{equation} is required. However, this requirement to the model parameters will not reduce the generality of our analysis so much because we are interested in the case that the typical length-scale for the modification of dispersion relation, $k_0^{-1}$, is sufficiently small. As in the case of $\delta$, we introduce an expansion parameter \begin{equation} \epsilon:=|\tilde\epsilon(x_2)|, \end{equation} and we neglect the terms that induce the relative error of $O(\epsilon^2)$ or smaller in the amplitude of the wave function. As for $\delta$, we also keep the terms up to linear order in $\delta$. Here one remark is in order. We imposed a further restriction $|x|>x_2$ to evaluate the explicit form of the solution (\ref{int2}). We stress, however, that the solution (\ref{int2}) itself is valid throughout the region $|x|<x_1$. To evaluate (\ref{int2}) by using the method of steepest descents, we need to know the value of $s$ at saddle points which are determined by solving \begin{eqnarray} f'(s) & \equiv & f'_0(s)+f'_1(s) \cr &=& \left[1+{\tilde\epsilon^{2}(sx)^{2}\over{2}} -\left(1+{i\omega \over{\kappa}}\right)(sx)^{-1}\right] +(sx)^{-1}W^{(1)}=0. \end{eqnarray} We solve this equation by assuming that the solution is given by a power series expansion with respect to $\delta$ as \begin{equation} s_{\pm } = s_{0\pm} + s_{1\pm}+ s_{2\pm}\cdots. \end{equation} For our present purpose, it is enough to find the solution in the form of series expansion with respect to $\epsilon$. One solution of $x s_0$ is of $O(1)$, and the integration along the path through this saddle point cannot be evaluated by the method of steepest descents. The other two solutions are given by \begin{equation} x s_{0\pm}:=\mp \left(\sqrt{-\tilde\epsilon^{2}\over 2}\right)^{-1} -{1+i\omega/\kappa \over 2} \pm \sqrt{{-\tilde\epsilon^{2}\over 2}} {3(1+i\omega/\kappa)^2\over8} +O(\epsilon^{2}), \label{sai} \end{equation} and \begin{eqnarray} xs_1 = &\pm& \left(\sqrt{-\epsilon^2\over 2}\right)^{-1} \left({\delta_0\over 4}-{\delta_1\over 4}+ \delta_2\right) +{1\over 4}\left[\left(3+{i\omega\over \kappa}\right)\delta_{0}+ \left(-3+{i\omega\over\kappa}\right)\delta_{1} +8\delta_{2}\right] \cr & \mp & {1\over 32}\sqrt{-\epsilon^2\over 2}\left(1+{i\omega\over \kappa}\right) \left[\left(9+{i\omega\over \kappa}\right)\delta_{0} -3\left(3-{5i\omega \over \kappa}\right)\delta_{1} +4\left(5-{3i\omega \over \kappa}\right)\delta_{2} \right]+O(\epsilon^{2}). \label{anten2p} \end{eqnarray} Since $ |s_{max} / s_{0\pm}| \sim \sqrt{1/\tilde \delta} \gg 1$, these saddle points are contained in the region in which the expansion with respect to $\delta$ is valid. In the following, to keep the notational simplicity, we abbreviate the subscript $\pm$ from $s_{0\pm}$ and $s_{1\pm}$ unless it causes any ambiguity. Now we evaluate the integration (\ref{int2}) by using the method of steepest descents. For the contour given in Fig.~1, only the saddle point $s_{+}$ dominantly contributes to the integration inside the horizon. For our present purpose, the formula \begin{equation} \psi(x) \approx {-\sqrt{2\pi}e^{xf(s_{\pm})}\over{(-xf''(s_{\pm}))^{1/2}}} \left[1-{5\over{24}}{(xf'''(s_{\pm}))^{2}\over{(xf''(s_{\pm}))^{3}}} +{1\over 8}{xf^{(4)}(s_{\pm})\over {(xf''(s_{\pm}))^{2}}} +O(\epsilon^2)\right], \label{anten3} \end{equation} is accurate enough to keep the correction up to $O(\epsilon^{1})$. The details of calculation to evaluate Eq.~(\ref{anten3}) up to $O(\epsilon, \delta)$ is given in Appendix \ref{appendixB}. In the end, we obtain \begin{eqnarray} \psi(x) &\approx & \sqrt{2 \pi \kappa} (k_0 \Omega_0)^{-{i \omega \over \kappa}-{1\over 2}} (-2 \kappa x)^{- {3 \over 4} - {i \omega \over {2 \kappa}}} \exp \left(-{2 \over 3} \sqrt{2\kappa}k_0 \Omega_0 (-x)^{3/2} +W + O(\epsilon^{2} , \delta^{2}) \right), \label{fsolution} \end{eqnarray} where \begin{eqnarray} W&=& -\left(\sqrt{-\tilde \epsilon^2\over 2}\right)^{-1}\left( -{1\over 10}\tilde\delta_0 +{1\over 10}\tilde\delta_1 -{2\over 5}\tilde\delta_2\right) \cr && +\left({i\omega \over {4 \kappa}}+{3\over8}\right)\tilde\delta_0 +\left({i\omega \over 4\kappa} - {3\over 8}\right)\tilde \delta_1 +{1\over 2}\tilde \delta_2 \cr && +\sqrt{{-\tilde\epsilon^{2} \over 2}} \Biggl[\left(-{41\over 48}-{i\omega\over \kappa} -{1\over 4}\left({i\omega\over \kappa}\right)^2\right) +\left({11\over 64}+{1\over 4}{i\omega\over \kappa} +{1\over 16}\left({i\omega\over \kappa}\right)^2\right)\tilde\delta_0 \cr && \quad\quad +\left(-{11\over 64}+{3\over 4}{i\omega\over \kappa} +{15\over 16}\left({i\omega\over \kappa}\right)^2\right)\tilde\delta_1 +\left(-{5\over 16}-{i\omega\over \kappa} -{3\over 4}\left({i\omega\over \kappa}\right)^2\right)\tilde\delta_2 \Biggr]. \label{name2} \end{eqnarray} We can immediately see that amplitude of the wave function reduces exponentially as we decrease $x$ (as we increase $-x$). \begin{figure}[tbh] \vspace*{0cm} \centerline{\epsfxsize5cm \epsfbox{himeron3.eps}} \caption{The deformed integration contour to evaluate $\psi(x)$ outside the horizon. In $x>0$ region, the saddle points move to the neighborhood of the imaginary axis. The contours $C_1$ and $C_2$ are chosen to pass through these saddle points and to be able to evaluate the integrations along them by the steepest descents. The remaining part of the integration contour which goes around the branch cut is called $C_3$. } \label{fig3} \end{figure} Next we turn to evaluate $ \psi (x) $ for $ x > 0 $. We use the method of steepest descents again. In the present case, the location of saddle points moves to points on the imaginary axis on the complex $s-$plane. The leading order approximation is given by \begin{equation} xs_{0 \pm} = \pm {\sqrt{2} i\over \epsilon_2} + \cdots. \label{sao} \end{equation} Therefore, to evaluate (\ref{int2}) by using the method of steepest descents, we need to deform the contour of integration. At this point, we must take account of the existence of a branch cut emanating from $s=0$, which originates from the log-term in the integrand. We choose this branch cut along the negative side of the real axis. Then, deforming the contour so as to go through these two saddle points we find that the contour is divided into three pieces as shown in Fig.~\ref{fig3}. We respectively denote them by $C_1 , C_2$ and $C_3$. $C_1$ and $C_2$ are the contours passing through the saddle points $s_-$ and $s_+$, respectively. Both contours have a new boundary point which is chosen to satisfy $|s|<s_{max}$. The contour $C_3$ connects these two newly introduced boundary points, going around the origin in the anti-clockwise manner. {}First we evaluate the integrations along the contours $C_1$ and $C_2$ by using the method of steepest descents. Just repeating the same calculation as in the case of $x<0$, these integrations are evaluated as \begin{equation} \psi_{1,2} (x) = e^{\mp \pi \omega / 2 \kappa }{ 1 \over { \sqrt{ k_0 \Omega_0} }} (k_0 \Omega_0)^{-{i \omega \over \kappa}} \sqrt{2 \pi \kappa} (2 \kappa x)^{ - {3 \over 4} - {i \omega \over {2 \kappa}}} \exp \left( \mp i {2 \over 3} \sqrt{2 \kappa} k_0 \Omega_0 x^{3/ 2} +W_{1,2} +O(\epsilon^{2},\delta^{2})\right), \label{near12} \end{equation} where \begin{eqnarray} W_{1,2}&=& \mp i {\sqrt{2} \over \tilde \epsilon}\left( -{1\over 10}\tilde\delta_0 +{1\over 10}\tilde\delta_1 -{2\over 5}\tilde\delta_2\right) \cr && +\left({i\omega \over {4 \kappa}}+{3\over8}\right)\tilde\delta_0 +\left({i\omega \over 4\kappa} - {3\over 8}\right)\tilde \delta_1 +{1\over 2}\tilde \delta_2 \cr && \mp i {\tilde\epsilon\over \sqrt{2}} \Biggl[\left(-{41\over 48}-{i\omega\over \kappa} -{1\over 4}\left({i\omega\over \kappa}\right)^2\right) +\left({11\over 64}+{1\over 4}{i\omega\over \kappa} +{1\over 16}\left({i\omega\over \kappa}\right)^2\right)\tilde\delta_0 \cr && \quad\quad +\left(-{11\over 64}+{3\over 4}{i\omega\over \kappa} +{15\over 16}\left({i\omega\over \kappa}\right)^2\right)\tilde\delta_1 +\left(-{5\over 16}-{i\omega\over \kappa} -{3\over 4}\left({i\omega\over \kappa}\right)^2\right)\tilde\delta_2 \Biggr]. \label{name3} \end{eqnarray} Next we consider the integral along the $C_3$ contour. Here, we divide $xf(s)$ given in Eq.~(\ref{xfs}) into two parts as \begin{eqnarray} && x \bar f_0(s):= xs + \left(-1-{i \omega \over \kappa} \right) \ln s \cr && x \bar f_1(s):= {\tilde\epsilon^2 (xs)^3\over 6}+xf_1(s), \label{newsep} \end{eqnarray} and we expand $e^{x\bar f_1(s)}$ assuming that $x\bar f_1(s)$ is small. From the validity of such an expansion, it is required that $|x \bar f_1(s)|\ll 1$. As for the case with large $|s|$, the integrand becomes exponentially small when $|xs|\gg 1$. There we do not have to mind at all even if $x\bar f_1(s)$ becomes large negative. In the restricted region satisfying $|xs|\alt 1$, it is easy to see that $x\bar f_1(s)\ll 1$ is always guaranteed. As for the case with small $|s|$, we do not have to consider the situation in which $|s|$ becomes extremely small because there is no requirement on the choice of contour except for being inside the saddle points. For example, if we choose the contour to be $|s|\agt 1$, $|x \bar f_1(s)| \ll 1$ is guaranteed. Therefore, we find that it is allowed to expand $e^{x\bar f_1(s)}$ as $1+x\bar f_1(s)+\cdots$. After this expansion, introducing a new variable $z$ by $e^{-i\pi}z:= sx$, the integration along $C_3$ is written as \begin{equation} \psi_{3} (x) \approx x^{i \omega \over \kappa} \int_{\bar C_{3}} \, dz \, (-z)^{-1 - {i \omega \over \kappa}} e^{-z}[1+x \bar f_1], \label{c3} \end{equation} where $\bar C_3$ is the contour in the complex $z-$plane corresponding to $C_3$. Since the integrand becomes exponentially small at the boundaries, we are allowed to continue the contour to $\infty$. Then, using the integral representation of a gamma function, the leading term corresponding to $1$ in the square brackets of Eq.~(\ref{c3}) is expressed as \begin{equation} \psi_{3} (x) \approx -2 \sinh (\pi \omega / \kappa ) \Gamma (-i \omega / \kappa) \, x^{i \omega / \kappa}. \label{gamma} \end{equation} Next, we consider the remaining terms in Eq.~(\ref{c3}). Let us express $x\bar f_1$ as $x\bar f_1=\sum a_n (-z)^n$, where the coefficients $a_n$ are non-dimensional constants. Then, by using the integral representation of the gamma function, we can evaluate the contribution from each term as \begin{equation} a_n \int_{\bar C_{3}} \, dz \, (-z)^{n-1-i\omega/\kappa} e^{-z} =-2 a_n \sinh\left(\pi\left({\omega\over\kappa}+in\right)\right) \Gamma\left(n-{i\omega\over \kappa}\right), \end{equation} and we find that its relative order is simply determined by the order of $a_n$. Hence, to find the expression correct up to $O(\epsilon^1, \delta^1)$, the only term that we must keep is \begin{equation} x \bar f_1(s)\approx-{i\omega \over {2\kappa}} \left({i\omega \over {\kappa}}+1 \right) \tilde \delta_1(sx)^{-1}. \end{equation} Thus, we finally obtain \begin{equation} \psi_{3} (x) = -2 \sinh (\pi \omega / \kappa ) \Gamma (-i \omega / \kappa) \, x^{i \omega / \kappa} \left(1-{i\omega \over {2\kappa}} \tilde \delta_1 +O(\epsilon^{2}, \delta^{2}) \right). \label{near3} \end{equation} In this section, we approximately solved Eq.~(\ref{horizon feq}) with the boundary condition that the wave function decreases exponentially inside the horizon. We evaluated the explicit form of the approximate solution in the region $x_1>x>x_2$ as \begin{equation} \psi(x) = \psi_{1}(x) + \psi_{2}(x) + \psi_{3}(x), \label{x>0} \end{equation} where each component $\psi_i(x)$ is given by Eq.~(\ref{near12}) or Eq.~(\ref{near3}). This expression is correct up to $O(\epsilon^1, \delta^1)$ . \subsection{the case far from the horizon} In the region far from the horizon, spacetime will become almost flat. In this region we assume that the rate of change of $1/\Omega(x)$ and $v(x)$ is sufficiently small. As we have seen for the asymptotic form of solutions in Sec.2, we have four independent solutions since the ODE~(\ref{ode}) is of fourth order. For the solutions with short wavelength corresponding to $k_{\pm}$, we can use WKB approximation to solve Eq.~(\ref{feq}). On the other hand, for the solutions with long wavelength corresponding to $k_{\pm s}$, we can solve Eq.~(\ref{feq}) perturbatively by treating the correction due to the modification of dispersion relation as small. To be strict, we restrict our consideration to the region $x>x_2$, where $x_2$ is that given in Eq.~(\ref{x_1}). In this region, we assume that the following relations \begin{equation} \Omega{d\over dx} {1\over \Omega},~ {1 \over v}{d\over dx} v,~ {1\over (1-v^2)}{d\over dx}(1-v^2) \alt {\omega\over 1-v^2}, \label{wkbyuukou} \end{equation} are satisfied. As for higher order differentiations, we also assume that they are all restricted like \begin{eqnarray*} \Omega {d^2\over dx^2}{1\over \Omega} \alt \left({\omega\over 1-v^2}\right)^2, \quad \Omega {d^3\over dx^3}{1\over \Omega} \alt \left({\omega\over 1-v^2}\right)^3, \cdots. \end{eqnarray*} By substituting the expansion (\ref{vomegat}), we find that these conditions are satisfied even in the region close to the horizon. Here, we define a quantity $\epsilon(x):= {\omega\over {k_0 \Omega(1-v^2)^{3/2}}}$, which reduces to $\tilde\epsilon(x)$ near the horizon. It will be natural to assume that $\epsilon(x)$ takes its largest value in the region close to the horizon, and hence $\epsilon(x)$ is at most of $O(\epsilon^1)$ owing to the restriction $x>x_2$. In the following, we construct approximate solutions valid up to $O(\epsilon^1)$ in the sense of $\delta\psi/\psi$. We begin with considering the solutions with short wavelength. Substituting the expression \begin{equation} \psi = e^{i\int^x dx'\, k(x')}, \label{WKB} \end{equation} into Eq.~(\ref{ode}), we write down the equation for $k(x)$. Neglecting the terms on which differentiations with respect to $x$ acted more than three times, \begin{eqnarray} {\left(1\over{{k_0}{\Omega}}\right)}^{2}k^{4} -(1-v^{2})k^{2}-2v{\omega}k+{\omega}^{2} & \approx & i{d\over{dx}}\left[2{\left(1\over{{k_0}{\Omega}}\right)}^{2}k^{3} -(1-v^{2})k-v{\omega}\right] \nonumber \\ & &+{3k'{}^2+4kk''\over k_0^2\Omega^2} +{12kk'\over k_0^2\Omega}\left({d\over dx}{1\over \Omega}\right) \nonumber \\ & & +{5 k^2\over 2k_0^2} \left(\left({d\over dx}{1\over \Omega}\right)^2 +{1\over\Omega}\left({d^2\over dx^2}{1\over \Omega}\right)\right), \label{feq2} \end{eqnarray} is obtained, where $~'~$ is used to represent a differentiation with respect to $x$. Denoting the left hand side of Eq.~(\ref{feq2}) by $F(k)$, we find that the first term on the right hand side is expressed as \begin{eqnarray*} {i\over 2}{d\over dx} \left({dF(k)\over dk}\right). \end{eqnarray*} We denote the remaining terms on the right hand side by $G(k)$. {}Following the standard prescription of the WKB approximation, the terms which contain differentiations with respect to $x$ is taken to be small. Accordingly, we also expand $k(x)$ in accordance with the number of differentiation as \begin{equation} k(x)=k^{(0)}(x)+k^{(1)}(x)+k^{(2)}(x)+\cdots. \label{g} \end{equation} After a slightly long but a straight forward calculation, we obtain \begin{eqnarray} k_{\pm} ^{(0)} &=& \pm k_0 \Omega \sqrt{1-v^2}\left(1\pm v\epsilon(x) -{1+2v^2\over 2}\epsilon^2(x)+O(\epsilon^3)\right)\cr & = & \pm k_0 \Omega \sqrt{1-v^2}+{v \omega \over 1-v^2} \mp {(1+2v^{2})\omega^{2} \over{2k_0 \Omega (1-v^{2})^{5/2}}} +\cdots, \label{k^0} \\ k_{\pm}^{(1)} & = &{i\over2}{d\over{dx}} {\ln}\left(\pm k_0 \Omega (1-v^{2})^{3/2} (1\pm 4v\epsilon(x)+O(\epsilon^2)) \right)\cr &=& {i\over2}{d\over{dx}} {\ln}\left(\pm k_0 \Omega (1-v^{2})^{3/2}+4vw+\cdots\right), \\ k_{\pm}^{(2)} &=& \pm k_0\Omega\sqrt{1-v^2}\Biggl\{ {\epsilon^2(x)\over 8\omega^2} \Biggl[41v^2 v'{}^2+(1-v^{2})\left( 14(vv')'+18 vv'\Omega\left({1\over \Omega}\right)'\right) \cr &&\hspace*{3cm} +(1-v^2)^2\left(4\Omega\left({1\over \Omega}\right)''+ \biggr[\Omega\left({1\over \Omega}\right)'\biggr]^2\right)\Biggr] +O(\epsilon^3)\Biggr\}. \end{eqnarray} Now we turn to the solutions with small absolute values, i.e., $k_{\pm s}$. In this case, we cannot use the WKB approximation because the wavelength is not necessarily short compared with the typical scale for the background quantities to change. However, for the model with the standard dispersion relation, we have exact solutions for the field equation $k_{\pm s}= \pm\omega/(1\pm v)$. We can use them as the leading order approximation, which is solutions when we neglect the terms related to the modification of dispersion relation. If we substitute $k_{\pm s}\approx \pm\omega/(1\pm v)$ into the neglected terms, we find that all of them have relative order higher than $\epsilon^2$. At first glance, one may think that the terms corresponding to the second and third terms in the square brackets in Eq.~(\ref{feq2}) give a correction of $O(\epsilon^0)$, but they mutually cancel out. As a result, the equation to determine the correction $\delta k_{\pm s}:=k_{\pm s}\mp\omega/(1\pm v)$ is obtained as \begin{equation} -i\partial_x (1-v^2)\delta k_{\pm s}\pm 2\omega \delta k_{\pm s}= {1\over 2k_0^2} e^{-i\int^x k_{\pm s}^{(0)}(x') dx'} \left[\partial_x^2{1\over \Omega}\partial_x{1\over \Omega}\partial_x +\partial_x{1\over \Omega}\partial_x{1\over \Omega}\partial_x^2\right] e^{i\int^x k_{\pm s}^{(0)}(x') dx'}=:H_{\pm}(x). \label{deltak} \end{equation} The right hand side consists of the terms related to the modification of dispersion relation, and they are small of $\omega k_{+s}^{(0)}\times O(\epsilon^2)$. Different from the case for short wavelength modes, the equation to determine the correction becomes a differential equation. Therefore, we can say that the correction stays of $O(\epsilon^2)$ only when we are interested in the behavior of the solution within a small region such as $x_1>x>x_2$. Once an extended region is concerned, there is no reason why the correction stays of $O(\epsilon^2)$. In fact, we need to know the behavior of the solution both at infinity and in the matching region $x_1>x>x_2$. In such a case, the correction much larger than $O(\epsilon^2)$ can appear as explained in detail in Appendix \ref{appendixC}. Nevertheless, the origin of this correction is the effect of scattering due to the modified dispersion relation. Even if the observed spectrum of the emitted particles deviates from the thermal one due to this effect, it is still possible to adopt the interpretation that the spectrum is modified by the scattering during the propagation to a distant observer though it was initially thermal. Hence, we think that this effect should be discussed separately from the present issue. However, to be precise, we consider the case that the condition (\ref{wkbyuukou}) replacing $\alt$ with $\ll$ is satisfied. This is the case when $\omega$ is sufficiently large or when the functions $v(x)$ and $1/\Omega(x)$ rapidly converge to some constants at $x\agt x_1$. In such cases, we can think of the first term on the left hand side of Eq.~(\ref{deltak}) as small. Then, solving Eq.~(\ref{deltak}) iteratively, we find that the correction stays of $O(\epsilon^2)$. Therefore, we obtain \begin{equation} k_{\pm s} \approx \pm{\omega \over 1\pm v}(1+O(\epsilon^{2})). \end{equation} Consequently, we find that the solutions which behave like $e^{ik(x\to\infty) x}$ at infinity are given by \begin{eqnarray} && {\psi}_{\pm}(x) = e^{i\int^x k_{\pm}(y) dy},\quad\quad \psi_{\pm s}(x) = e^{i\int^x k_{\pm s}(y) dy}, \label{formalwkb} \end{eqnarray} where the integral constants are chosen appropriately. Here we recall that what we wish to know is not $k(x)$ but $\int^x dy\, k(y)$. Although we are keeping track of the error in the expression of $k(x)$, we cannot evaluate the error in the integral $\int^x_\infty k(y) dy$ when it is integrated from $\infty$ to the matching region where $x_1>x>x_2$. To overcome this difficulty, we need to make use of the existence of a conserved current \begin{equation} j=A(k(x)) e^{-2\int^x k_{(I)}(y) dy}, \end{equation} where \begin{eqnarray} A(k(x))=&&\omega v+(1-v^2)k_{(R)} \cr &&-{1\over k_0^2\Omega^2}\left\{2k_{(R)}(k_{(R)}^2-k_{(I)}^2) +4k_{(R)}\partial_x k_{(I)}+2k_{(I)}\partial_x k_{(R)} -\partial_x^2 k_{(R)}\right\}\cr &&+{2\over k_0^2}\left({1\over\Omega}\left({1\over\Omega}\right)'\right) \left(-2k_{(R)}k_{(I)}+\partial_x k_{(R)}\right) +{1\over 2k_0^2}\left({1\over\Omega}\left({1\over\Omega}\right)'\right)' k_{(R)}, \end{eqnarray} and $k_{(R)}$ and $k_{(I)}$ are the real and the imaginary parts of $k$, respectively. The derivation of $j$ is given in Appendix \ref{appendixD}. We evaluate this conserved current $j$ at $x\to\infty$, where all terms that contain differentiations with respect to $x$ vanish there. Adopting the normalization $\vert\phi\vert^2 = 1$ at $x\to\infty$, $j$ is determined as \begin{equation} j=\omega v_{\infty}+(1-v_{\infty}^2)k_{\infty} -{2\over k_0^2\Omega_{\infty}^2} k_{\infty}^3. \end{equation} For $l=\pm,\pm s$, by substituting $k_{l\infty}$ into the expression of $j$ in place of $k_{\infty}$, we also define the conserved current $j_l$ corresponding to $k_l$. Owing to the conservation of $j$, \begin{equation} \psi_l(x)=\sqrt{j_l\over A(k_l(x))} e^{i\int^x k_{l(R)}(y) dy}. \label{hozonpsi} \end{equation} By using this improved expression, we can calculate the explicit form of $\psi_l(x)$ in the region $x_1>x>x_2$ without any ambiguity except for the constant phase factor that does not alter the absolute magnitude of the wave function. Expanding the expression (\ref{hozonpsi}) in powers of $\tilde \delta_i$ with the substitution of Eqs.~(\ref{vomegat}), we evaluate $\psi_l(x)$ keeping the terms up to $O(\delta^1)$. Here in evaluating $A(x)$, the terms that include $k_{(I)}$ becomes higher order in $\epsilon$, and hence we can neglect them all. Consequently, we obtain \begin{eqnarray} && {\psi_{\pm}(x)\over \sqrt{j_{\pm}}} ={e^{i\alpha_{\pm}}\over {\sqrt{k_0 \Omega_0}}} (2 \kappa x)^{-{3\over 4}-{i\omega \over {2\kappa}}} \exp \left( \pm i {2\over 3} k_0 \Omega_0 \sqrt{2 \kappa} x^{3\over 2} +W_{\pm}+O(\epsilon^{2}, \delta^{2})\right), \cr && {\psi_{+s}(x)\over\sqrt{j_{+ s}}} = e^{i\alpha_{+s}} x^{i\omega \over \kappa} \exp\left(-{i\omega \over {2 \kappa}} \tilde \delta_1 +O(\epsilon^2,\delta^2)\right), \label{fwkb} \end{eqnarray} where $W_+$ and $W_-$ are no different from $W_2$ and $W_1$ in Eq.~(\ref{name3}), respectively. As noted above, there appears an integration constant $\alpha_l$ that cannot be determined by the present analysis, but it is guaranteed to be a real number. Here, we did not give the explicit form of $\psi_{-s}$ because we do not use it later. By comparing (\ref{x>0}) with (\ref{fwkb}), we find that the solution (\ref{x>0}) obtained in the region close to the horizon is matched to the solutions obtained in the outer region like \begin{equation} \psi(x)=\sqrt{2 \pi \kappa} \left( e^{-{\pi \omega \over{2 \kappa}}+i\alpha'_-} {\psi_-\over \sqrt{j_-}}+ e^{{\pi \omega \over 2 \kappa}+i\alpha'_+} {\psi_+\over \sqrt{j_+}} \right) - 2 \sinh \left( {\pi \omega \over \kappa} \right) \Gamma \left(-{i\omega \over \kappa} \right) e^{i\alpha_{+s}}{\psi_{+s}\over \sqrt{j_{+s}}}, \label{fsolution} \end{equation} where $\alpha'_{\pm}$ are also real constants. From this expression, we can read the coefficients $c_{+s},c_{-}$ as \begin{eqnarray} && c_{+s}=- 2 \sinh \left( {\pi \omega \over \kappa} \right) \Gamma \left(-{i\omega \over \kappa} \right) {e^{i\alpha_{+s}}\over \sqrt{j_{+s}}}(1+O(\epsilon^2,\delta^2)) \cr && c_{-}=\sqrt{2 \pi \kappa}e^{-{\pi \omega \over{2 \kappa}}} {e^{i\alpha'_{-}}\over \sqrt{j_{-}}} \left( 1 - {4v\omega\over{k_0\Omega}}(1-v^{2})^{-3/2}\right)^{-1/2} (1+O(\epsilon^2,\delta^2)). \end{eqnarray} The factor $\omega' v_g$ appears in the formula (\ref{number}) for the expectation value of the created particles. By differentiating the dispersion relation at infinity (\ref{disp}) with respect to $k_{\infty}$, this factor is easily calculated as \begin{equation} \omega'(k_{\infty}) v_g(k_{\infty}) :=(\omega-v k_{\infty}) {d\omega(k_\infty)\over dk_{\infty}} =v_{\infty}\omega +(1-v_{\infty}^2)k_{\infty} -{2k_{\infty}^3\over k_0^2\Omega_{\infty}^2}, \end{equation} and we find it to coincide with the conserved current. Thus, by considering the combination of $\omega'(k_{+s}) v_g(k_{+s})\vert c_{+s}\vert^2$ the factor $j_{+s}$ in $c_{+s}$ cancels, and the same is also true for $k_-$. Finally the expectation value of the number of created particle is evaluated as \begin{equation} N(\omega)={ 1 \over {e^{2 \pi \omega / \kappa}-1}} (1+O(\epsilon^{2},\delta^{2})). \end{equation} \section{analytic and numerical studies of the deviation from hawking spectrum} \vspace*{5mm} {\renewcommand{\arraystretch}{1.6} \begin{center} \begin{tabular}{c||c|c|c|c} $~~~~~~$ & $~~\epsilon^{-1}~~$ & $~~\epsilon^{0}~~$ & $~~\epsilon^{1}~~$ & $~~\epsilon^{2}~~$ \\ \hline &&&&\\[-6.1mm] \hline {$\delta^{0}$} & $x^{3/2}$ & $x^0$ & $x^{-3/2}$ & $x^{-3} $ \\ \hline $\delta^{1}$ & $x^{5/2}$ & $x^{1}$ & $x^{-1/2}$ & $x^{-2}$ \\ \hline $\delta^{2}$ & $x^{7/2}$ & $x^{2}$ & $x^{1/2}$ & $x^{-1}$ \\ \hline $\delta^{3}$ & $x^{9/2}$ & $x^{3}$ & $x^{3/2}$ & $x^{0}$ \end{tabular} \vspace*{5mm} {\small TABLE I. Table to explain the matching procedure.} \end{center} } In the preceding section, to obtain the flux of the created particles observed in the asymptotic region, where $v(x)$ is essentially constant, we propagated the near-horizon solution (\ref{x>0}), which satisfies the appropriate boundary condition, to the infinity by matching it with the outer-region solutions (\ref{fwkb}), which are valid in the region distant from the horizon. As a result, we could determine the coefficients $c_{l}$ and we found that the thermal spectrum is reproduced up to $O(\epsilon^1, \delta^1)$. To explain the matching procedure in more detail, here we present a table. In the construction of the near-horizon solution, the equation to be solved was expanded with respect to $\delta$, and we obtained an equation which correctly determines the terms up to $O(\delta^1)$. These terms correspond to the first two lines in the above table. Then, expanding them with respect to $\epsilon$ by restricting our consideration to the region $x_2<x<x_1$, we obtained the expression (\ref{x>0}) with (\ref{near12}) and (\ref{near3}), which contains the terms corresponding to the first $2\times 3$ elements in the table. On the other hand, in the region distant from the horizon, we first considered an expansion of the solution with respect to $\epsilon$, and calculated the corrections up to $O(\epsilon^1)$. Namely, the terms corresponding to the first three columns in the above table are obtained. As the next step, in the region of $x_2<x<x_1$, we expanded this expression also with respect to $\delta$ up to $O(\delta^1)$ by substituting (\ref{vomegat}) into Eq.~(\ref{name3}), and we obtained the first $2\times 3$ elements in the table. The expressions that we finally obtained are the outer-region solutions~(\ref{fwkb}). Such two fold expansions in both schemes are simultaneously valid only in the region $x_1<x<x_2$, where both $\tilde \epsilon(x)$ and $\tilde \delta_i(x)$ are small. As mentioned above, as long as $k_0$ is taken to be sufficiently large, this overlapping region always exists. Since both expressions obtained by using the above two different schemes are approximate solutions of the same equation, they must be identical if we take an appropriate superposition of the four independent solutions. In fact, we found that the near-horizon solution (\ref{x>0}) can be written as a superposition of the outer-region solutions as given in Eq.~(\ref{fsolution}). Now, let us look at the above table again. For each element in the table, we have assigned a power of $x$ that the corresponding terms possess. As we mentioned above, we need to choose an appropriate superposition of the four independent outer-region solutions to achieve a successful matching. The coefficients which determine the weight of this superposition is nothing but $c_l$. Now we should note that the condition to determine the coefficients $c_l$ will be completely supplied by matching the $x$-independent elements. Once these coefficients are determined, the agreement of the other $x$-dependent terms must be automatic from the consistency. The leading order $x$-independent elements consists of the terms of $O(\epsilon^0\delta^0)$, and it is easy to see that the second lowest one consists of the terms of $O(\epsilon^{2} \delta^{3})$. This fact tells us that the possible modification of the coefficients $c_{+s}$ and $c_-$ is at most of $O(\epsilon^{2} \delta^{3})$, and hence the possible deviation from the thermal radiation starts only from this order. Thus we find \begin{equation} N(\omega)={ 1 \over {e^{2 \pi \omega / \kappa}-1}} (1+O(\epsilon^{2} \delta^{3})). \label{thermal} \end{equation} Next we investigate the deviation from thermal radiation $\delta N/N:=(N-N_{thermal})/N_{thermal}$ in more detail. There are three different quantities of $O(\delta)$ as given in (\ref{ddelta}). We write them down as \begin{equation} \tilde\delta_0=x v'|_{x=0} =\kappa x, \quad \tilde\delta_1=x \kappa^{-1} v''|_{x=0} =:(\kappa x)b_1, \quad \tilde\delta_2=x \Omega_0 \left({1\over \Omega}\right)'_{x=0} =:(\kappa x)b_2, \end{equation} where we introduced non-dimensional model parameters $b_1$ and $b_2$. We also have various quantities of $O(\delta^2)$ and of $O(\delta^3)$ consisted of higher derivatives of $v$ and $1/\Omega$. They are \begin{equation} x^2 \kappa^{-1} v'''|_{x=0} =:(\kappa x)^2 b_3, \quad\quad x^2 \Omega_0 \left({1\over \Omega}\right)'_{x=0} =:(\kappa x)^2 b_4, \end{equation} and \begin{equation} x^3 \kappa^{-1} v^{(4)}|_{x=0} =:(\kappa x)^3 b_5, \quad\quad x^3 \Omega_0 \left({1\over \Omega}\right)''_{x=0} =:(\kappa x)^3 b_6, \end{equation} respectively. Also, $b_3$, $b_4$, $b_5$ and $b_6$ are non-dimensional model parameters. One may suspect that terms including factors proportional to $\kappa^{-2}$ such as $x \kappa^{-2} v'''|_{x=0}$ might appear among the correction terms of $O(\epsilon^2\delta^3)$. However, by repeating the same calculation that was given in Sec.4 with extra higher order derivative terms, we can verify that such factors do not appear. From this notion, we can expect that the deviation from the thermal spectrum is given by \begin{eqnarray} {\delta N\over N} ={\kappa^2\over k_0^2\Omega_0^2}\Biggl[ && \left(a_{000}+a_{001} b_1+a_{002} b_2 +(\mbox{7 other terms})\right) \cr && +\left(a_{03}+ a_{13} b_1+ a_{23} b_2\right) b_{3} +\left(a_{04}+ a_{14} b_1+ a_{24} b_2\right) b_{4} +a_{5} b_{5} +a_{6} b_{6}\Biggr] +O(\epsilon^4\delta^{6}), \label{dev} \end{eqnarray} where ``$a$''s are some functions of $\omega/\kappa$ which are independent of model parameters. Now, we numerically confirm that the deviation actually starts from this order. The following results are obtained by using {\it MATHEMATICA}. As an example, let us consider a model given by \begin{eqnarray} v & = & -{1\over 2}e^{-2 x - 3 x^2} -{1\over 2}, \label{vmodel} \\ \Omega & = & 9 e^{-x-x^2/2}+1. \label{Omegamodel} \end{eqnarray} For this model, we have $\kappa=1$, $b_1=1$ and $b_2=9/10$, and the other parameters also do not vanish. For this fixed model, we numerically calculated the deviation from the thermal spectrum for various values of $1/k_0^{2}$. The frequency $\omega$ was fixed to 1 since our main interest is in the modes whose observed frequency at infinity becomes comparable with Hawking temperature ($= \kappa /2 \pi$). The results of the numerical calculation are shown in Fig.3 by the filed circles. The horizontal axis is $\log (1/k_0)$ and the vertical one is $\log \delta N/N$. The data points are fitted well by a liner function (the solid line) with its gradient, $1.99742$, which perfectly agrees with the expectation represented by Eq.(\ref{dev}). At this point, one may notice that the deviation we obtained here is much larger than that given by Corley and Jacobson\cite{JacCor}, in which a model with \begin{equation} v = {1\over 2} \{\tanh [(2\kappa x)^{2}]\}^{1/2}-1, \end{equation} and $\Omega\equiv 1$ was considered \footnote{T. Jacobson suggested us the existence of this discrepancy.}. The outstanding feature of their model is that $b_1=b_2=b_3=b_4=b_5=b_6=0$. Hence, the terms of $O(\epsilon^2\delta^3)$ in Eq.(\ref{dev}) reduce to $a_{000}{\kappa^2 /k_0^2\Omega_0^2}$. If $a_{000}\equiv 0$, all the terms of $O(\epsilon^2\delta^3)$ in the deviation $\delta N/N$ disappear, and it turns out to be $O(\epsilon^4\delta^6)$. If so, the discrepancy between two calculations can be understood. To show that this is certainly the case, we repeated the numerical calculation for the same model that was discussed in Ref.\cite{JacCor}. The resulting $\delta N/N$ calculated for various values of $1/k_0$ were also plotted in Fig.3 by the open squares. Again, the data points in logarithmic plot are fitted well by a linear function. But, this time, its gradient is 4.06935, which indicates that the deviation is actually caused by the terms of $O(\epsilon^4\delta^6)$. Now, we can conclude that $a_{000}\equiv 0$. Although this result might be interesting, we do not pursue this direction of study in this paper. Here, we would like to focus on another interesting aspect that is anticipated by the expression (\ref{dev}). With moderate values of model parameters, the deviation $\delta N/N$ stays small for a sufficiently large $k_0$. However, conversely, we can expect that the deviation from the thermal spectrum becomes large if we consider some extreme modifications of the special reference frame. Especially, when we consider the limiting case in which $\Omega_0^2\to 0$, the expression (\ref{dev}) diverges. Although the approximation used to obtain the analytic expression (\ref{dev}) is no longer valid in this limit, we can still expect that the resulting spectrum will significantly differ from the thermal one. As we mentioned below Eq.(\ref{newv}), there is a lower bound on $\Omega^2(x)$. The possible smallest value of $\Omega^2(x)$ is $1-\tilde v^2$, which is realized when $v(x)=0$. Hence, we find $v(x)\approx 0$ near $x=0$ in this limiting case. Recall that $v(x)$ was the coordinate velocity of the integration curves of $u$. Hence, vanishing $v(x)$ means that we adopt the reference frame corresponding to the static observers. Here, we present the results of our numerical calculation, which shows that the deviation $\delta N/N$ can be large for some cases. Since we also want to demonstrate that the drastic change of spectrum can occur just in the consequence of the change of the special reference frame, we varies only the function $\gamma(\tilde x)$, which was defined in Eq.(\ref{tconst}). The model of $\tilde v(\tilde x)$ is kept unchanged. As for a model of $\tilde v(\tilde x)$, we assume the same form that is given in Eq.(\ref{vmodel}). As for $\gamma(\tilde x)$, we adopt \begin{equation} \gamma={-\tilde v-(1/2)\over \rho^{-1} -\tilde v}, \end{equation} with $\rho \in [0,1)$. $\rho=0$ corresponds to the original model associated with the freely falling observers, and $\rho=1$ corresponds to the case with $\Omega_0^2=0$. With this choice of $\gamma(\tilde x)$, the following two conditions are satisfied. One is that $v(x)$ stays negative for all positive $x$. The other is that $\Omega_{\infty}^2= 1$. We calculated the deviation $\delta N/N$ for various values of $\rho$, and the results were shown in Fig.4. As was expected, the deviation becomes large for small $\Omega_0^2$. This plot raises an interesting speculation that $N$ might converge to $0$ in the $\Omega_0^2\to 0$ limit. Although we have not confirmed it yet, it is very likely that this is the case because the situation in this limit is very similar to the case in which we set a static mirror surrounding the event horizon of the black hole. The calculation for small $\Omega_0^{2}$ were tried, but it was found to be out of validity of our present computation code. Anyway, we conclude that, even if $\kappa/k_0$ is sufficiently small, the deviation from the thermal spectrum can be large if the combination $\kappa/k_0\Omega_0$ becomes large. It should be noted that the integration curves of $u$ are not required to be suffered from infinite acceleration to achieve such a small but non-zero $\Omega_0^2$. \begin{figure}[tbh] \vspace*{0cm} \centerline{\epsfxsize7cm \epsfbox{fig1.eps}} \caption{The logarithmic plot of $\delta N/N$ as a function of $1/k_0$ for two different choice of $v(x)$. The filled circles(our model) and the open squares(model in Ref.[5]) represent the numerical data points for respective models. Each solid line corresponds to a linear function which fits the data points. } \label{fig4} \end{figure} \begin{figure}[tbh] \vspace*{0cm} \centerline{\epsfxsize7cm \epsfbox{fig2.eps}} \caption{A plot of $\delta N/N$ as a function of $1/\Omega_0$ for the model given by Eq.(5.6). } \label{fig5} \end{figure} \section{Conclusion} We studied the particle creation in a model which is a generalization of the Unruh's toy model. In his model, the field equation for a scalar field is modified by introducing a non-standard dispersion relation. To do so, we necessarily violates the Lorentz invariance. This radical change of theory is originally motivated by the possible existence of the effect due to the unknown physics at Planck scale. However, as is explained in Introduction, there is another point of view, on which it is also meaningful to study this model as an effective theory which takes into account the interaction between various fields even if we believe that the Lorentz invariance is exact, In the original model, the dispersion relation is modified on the basis of the freely falling observers. In our present work, we generalized the choice of the reference frame with respect to which we set the non-standard dispersion relation. Extending the analytic method developed by Corley~\cite{Corley}, we have shown that the thermal spectrum of radiation from a black hole is almost reproduced as long as the modification of the special reference frame is not too extreme. In this analysis, we assumed $\omega\approx \kappa$, where $\omega$ is the frequency of the emitted photon observed at the spatial infinity and $\kappa$ is the surface gravity of the black hole. We have also obtained a strong suggesting that the deviation from the exact thermal spectrum appears from $O(\kappa^2/k_0^2)$, where $k_0$ is the typical wave number corresponding to the modification of the dispersion relation. This speculation has been confirmed numerically. Of course, we should not stress this small deviation from the Hawking spectrum. In the ordinary model with the Lorentz invariant dispersion relation, the thermal radiation at the temperature $T$ for the static observer is observed as the thermal radiation at the temperature $\sqrt{1-\beta\over 1+\beta} T$ for the observer moving with the radial velocity $\beta$. This argument holds in general whatever the source of the outward pointing radiation is because it is a direct consequence of $\omega=k$. However, in the present modified model, the Lorentz invariance is violated from the beginning. We can easily see that $(\omega-k)/\omega$ is also of $O(\omega^2/k_0^2)$. Hence, even if the exact Hawking spectrum is reproduced for one specific free-falling observer, it cannot be so for the other free-falling observers. On the other hand, the result that we obtained analytically also suggests that the deviation from the thermal spectrum can be large if we consider some extreme situations. With the aid of numerical methods, we also examined one of such extreme situations. We considered a sequence of different special reference frames which ranges from the case in which the observers associated with the special reference frame are freely falling into black hole to the case in which they are kept from falling into it. We found that, in the latter limiting case, the spectrum of radiation can significantly differ from the thermal one, even though $\omega^2/k_0^2$ is small. It will be important to study the physical meaning of this result. But, since the central issue of this paper is to develop the analytic treatment of our new model, we have not performed detailed numerical studies yet. We will return to this issue in future publications. \acknowledgements {We would like to thank Misao Sasaki for useful suggestions and comments. Y.H. also thanks Fumio Takahara for his continuous encouragement and Uchida Gen for the useful conversation with him. Additionally, he thanks Shiho Kobayashi for his appropriate advice. Lastly we wish to thank Ted Jacobson for fruitful discussions in YKIS'99. }
\section{Introduction} The last step in comparing theoretical isochrones to globular cluster colour-magnitude diagrams (CMD) is the transformation of luminosity and effective temperature into brightness and colour. While there are a number of both theoretical and empirical sets of transformations available, this last step is, nonetheless, difficult and critical. Known shortcomings in the theoretical transformations which are obtained from theoretical non-grey model atmospheres calculations are, e.g., incomplete line lists for the opacity computation which affect the derived spectra and broad-band colours, the treatment of atmospheric convection which affects the colour determination as well (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{gcc:96}), and an incomplete coverage of the $\log g$--$\log T_{\rm eff}$ and composition parameter space, while empirical relations are naturally based on a limited number of stars. The transformations of Kurucz and coworkers (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{kur:92}; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{cgk:97}) have well-known problems with the colours of red giants (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{gcc:96}; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sdw:97}; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{gva:98}), but are nevertheless widely used. It is generally believed that the bolometric corrections, which are more important for most methods of cluster age determinations, are reliable, at least in their differential properties, and indeed results based on different transformations are in reasonable agreement (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sdw:97}). As long as there are no unified stellar models available, which treat the stellar interior and the non-grey atmosphere together in one complete and realistic model (but see \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{bern:98} for a first unified solar model), such transformations are the only way to produce observable photometric quantities. There is also the persistent concern about the unsolved problem of superadiabatic convection in the envelopes of cool stars, which determines surface temperature and thus colour. In our own calculations we have used the standard mixing-length theory with one constant value for the mixing-length parameter, because even with this very simplistic choice we are able to match both the solar radius and the effective temperature of metal-poor giants (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sw:98}). The existence of a rather constant value for this parameter is also supported by comparing stellar models with the structure of convective envelopes computed by means of 2-dimensional hydrodynamical models (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{lfs:97}; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{frey:98}). However, there are other convection theories possible (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{cm:91}), which lead to different $T_{\rm eff}$ along an isochrone compared to the standard mixing-length approach. The discussion in this paper is therefore restricted to our own treatment of convection, and the conclusions could be different for a different choice. The confidence in ages derived from theoretical isochrones is higher if the isochrone is able to reproduce the complete CMD from the lower main sequence (MS) to the tip of the red giant branch (RGB) to high accuracy, even if in practice only single points along the isochrone may be used for determining the cluster age (e.g.\ turn-off and horizontal branch). Reliable colours become crucial for methods that use colour differences, e.g.\ that between turn-off (TO) and RGB, for (absolute) age determinations, for determining distance and age from the MS-fitting technique, or for deriving correct integrated colours from theoretical isochrones, which constitutes an necessary step for studying old stellar populations in unresolved galaxies. In our previous papers on this subject (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sdw:97}; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sw:97}, ``SW97''; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sw:98}, ``SW98'') we have taken great care to find transformations from the theoretical Hertzsprung-Russell-Diagram (HRD) to the CMD in $V-(B-V)$, which give a satisfying fit for {\em all} metallicities ranging now from ${\rm [Fe/H]} = -2.3$ to $-0.6$. We found that a combination of the transformations of \def\astroncite##1##2{##1 (##2)}\@internalcite{bk:78} and \def\astroncite##1##2{##1 (##2)}\@internalcite{bk:92} (with an appropriate colour shift of the latter ones to enforce continuity in (B-V) at the common temperature of 6000 K) satisfied our needs. Work on globular cluster (GC) dating during the last few years has received new attention. Due to improved models and methods and lately due to HIPPARCOS-based distances to clusters, much lower ages are obtained than before, such that, at this time, no serious conflict between cluster ages and the age of the universe exists (SW98; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{chab:98}). In order to extend our work to a larger sample of galactic halo clusters and in particular to those of the bulge (both needed for questions about the formation of the Galaxy), it is necessary to work in the $V-(V-I)$--CMD\footnote{If not stated differently, the $I$-band refers to the Cousins system}, because there is now an increasing number of high-quality observations of clusters in $V-(V-I)$. However, up to now, well-tested isochrones for this colour have not been available. In this paper, we demonstrate that existing empirical and theoretical transformations provide very different $T_{\rm eff}-(V-I)$ relationships with only very few being in agreement with each other. For the choice as to which transformation we apply to our isochrones, we want to see the following requirements being fulfilled: (i) the transformation should give RGB colours consistent with the determined age; (ii) an isochrone that fits a cluster well in $(B-V)$, should do so in $(V-I)$, too, for E(V-I) values compatible with the corresponding E(B-V). In Sect.~2 we will present a number of recently published theoretical and empirical transformations and apply them to our theoretical isochrones. We will show that considerable differences in the resulting colours exist and that a priori there is no preferred transformation to be selected. However, at the same time the reliability of the bolometric correction in $V$ will be reconfirmed. We will provide arguments for our selection of transformations for MS and RGB. In Sect.~3, we will then present tests verifying its usefulness for the case of our own isochrones applied to globular clusters of various metallicities. By no means this rather pragmatic approach is what one really wants for isochrone fittings, but at present it appears to be the only approach promising convincing results. Finally, a discussion and summary closes the paper (Sect.~4). \section{A comparison of different published colour-transformations} \subsection{Transformation sources} For the purpose of this investigation we have used a number of recently published transformations, which we briefly describe in the following: --\def\astroncite##1##2{##1}\@internalcite{bk:78} (\def\astroncite##1##2{##2}\@internalcite{bk:78}; \def\astroncite##1##2{##2}\@internalcite{bk:92}, ``BK92''): This is the source for synthetic colours we have been using in our previous papers, showing that an appropriate combination (henceforth called ``BK'') yields satisfactory CMD-fits for all metallicities in the $V-(B-V)$ plane. The transformations are based on theoretical atmospheres and are available for all metallicities and both for dwarfs and giants. $(V-I)$ colours are available only for the cooler stars on the lower MS and on the RGB (BK92). --ATLAS9: This is a set of spectra and colours obtained from a model atmosphere grid computed with the ATLAS9 code (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{kur:92}), covering the range of O-K stars. The first set of ATLAS9 colours was discussed in \def\astroncite##1##2{##1 (##2)}\@internalcite{kur:92}. The convective models of this set were later revised (``K95''; see \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{bcp:98}); the influence of the treatment of convection in the ATLAS9 model atmospheres on some colour indices has been discussed by \def\astroncite##1##2{##1 (##2)}\@internalcite{cgk:97}. The latest grid of ATLAS9 colours, computed from ``no-overshoot'' models is presented in \def\astroncite##1##2{##1 (##2)}\@internalcite{bcp:98} for the solar metallicity. In this paper we will use this latest set of colours (hereinafter ``BCP98'') extended to lower metallicities by Castelli (1997, private communication). --\def\astroncite##1##2{##1}\@internalcite{lcb:97} (\def\astroncite##1##2{##2}\@internalcite{lcb:97}, \def\astroncite##1##2{##2}\@internalcite{lcb:98}; ``LCB''): Synthetic spectra and colours based on various sets of theoretical atmospheres, among them K95 (covering almost completely the range of gravities and $T_{\rm eff}$ spanned by our isochrones). The theoretical spectra for the solar metallicity have been corrected to match empirical colour-$T_{\rm eff}$ relations; since comprehensive empirical calibration data have only been available for the full temperature sequences of solar-abund\-ance giant and dwarf stars, the spectra (and derived colours) of lower metallicity models have been consequently corrected in such a way as to preserve the monochromatic flux ratios predicted by the models. -- \def\astroncite##1##2{##1}\@internalcite{hab:99} (\def\astroncite##1##2{##2}\@internalcite{hab:99}; ``Next-Gen''): Grid of theoretical model atmospheres, colours and bolometric corrections covering the main sequence of globular clusters, with $T_{\rm eff}$ ranging between 3000 and 10000 K, $\log g$ between 3.5 and 5.5 and [M/H] between -4.0 and 0.0. The atomic line list is the same as in \def\astroncite##1##2{##1 (##2)}\@internalcite{kur:92}, but the treatment of molecular, bound-free and free-free opacity sources, as well as the equation of state and the selected solar metal distribution are different. For convection, standard mixing-length theory with a constant parameter of 1.0 was used. Note that all theoretical work up to now is for solar metal ratios only, i.e.\ no enhancement of the $\alpha$-elements has been considered. Work along this line is in progress (Castelli 1999, in preparation). --\def\astroncite##1##2{##1}\@internalcite{aam:96} (\def\astroncite##1##2{##2}\@internalcite{aam:96}; ``AAM96''): Empirical determination of the $T_{\rm eff}$-colours relation for main sequence stars as a function of metallicity, using a large sample of dwarfs and subdwarfs. Since AAM96 use the Johnson--$I$-band, we have transformed Johnson--$(V-I)$ into Johnson--Cousin--$(V-I)$ colours according to \def\astroncite##1##2{##1 (##2)}\@internalcite{fer:83} throughout this paper. --\def\astroncite##1##2{##1}\@internalcite{mfo:98} (\def\astroncite##1##2{##2}\@internalcite{mfo:98}; ``M98''): Empirical determination of bolometric correction-colour, colour-colour and $(V-K)-T_{\rm eff}$ relations obtained from photometric observations of 6500 GC giants covering metallicities from ${\rm [Fe/H]} = -2.2$ to $0.0$. The results are averaged over two mean metallicity ranges below (``poor'') and above (``rich'') ${\rm [Fe/H]}= -1.0$. --\def\astroncite##1##2{##1}\@internalcite{bcm:98} (\def\astroncite##1##2{##2}\@internalcite{bcm:98}; ``BCM98''): Empirical colour--$(V-K)$ relations for giants in GC of low and intermediate metallicity as well as one open cluster. These relations, together with an empirical $T_{\rm eff}-(V-K)$ relation taken from the literature, provide direct $T_{\rm eff}$-colour transformations for GC giants, taking into account their dependence on metallicity. In the empirical work, $\alpha$-element enhancement is considered only as far as the observed stars show it. There is no systematic differentiation between solar-type and $\alpha$-enhanced populations. \subsection{Isochrones in the $V-(B-V)$--CMD and the bolometric correction} The theoretical isochrones used throughout this paper were computed for metallicities ranging from ${\rm [Fe/H]}=-2.3$ up to ${\rm [Fe/H]}=-0.6$, and are based on the same stellar evolutionary tracks as in our previous papers (e.g.\ SW98), except for the fact that meanwhile we have extended all the calculations to the tip of the RGB, i.e.\ to the onset of the core helium flash for all masses considered. The ZAHB models have been calculated from RGB-tip models. All isochrones\footnote{The isochrones are available on request from the authors.} used here are for $\alpha$-element enhanced mixtures ($\langle{\rm [\alpha/Fe]}\rangle$=0.4). \begin{figure} \centerline{\includegraphics[draft=false,scale=0.60]{figBC.ps}} \caption[]{Comparison of $BC_{V}$ from BK with empirical data by \def\astroncite##1##2{##1 (##2)}\@internalcite{fpc:81} (see text). The outlier in panel c is in the original data} \protect\label{figBC} \end{figure} As previously mentioned, our source for $(B-V)$ colours and bolometric corrections ($BC_{V}$) is BK. In the previous papers (e.g.\ SW98) we have shown that by using these colours we were able to obtain a good fit to all clusters sequences covered by our isochrones, to derive GC reddenings in good agreement with the results from \def\astroncite##1##2{##1 (##2)}\@internalcite{z:85}, and to reproduce the HIPPARCOS subdwarf colours given by \def\astroncite##1##2{##1 (##2)}\@internalcite{gfc:97}. As for the $BC_{V}$ scale, the zero point was calibrated to the empirical scale by \def\astroncite##1##2{##1}\@internalcite{aam:95} (\def\astroncite##1##2{##2}\@internalcite{aam:95}, \def\astroncite##1##2{##2}\@internalcite{aam:96}) for metal-poor MS stars; it reproduces the empirical $BC_{V}$-values for the whole range of metallicities and MS temperatures covered by the isochrones within $\pm$0.02 mag. We required in particular that the solar $V$ magnitude ($V_\odot$ = 4.82) is reproduced. With all the isochrones extended now up to the RGB-tip, we can also compare our $BC_{V}$ scale for RGB stars with the empirical results by \def\astroncite##1##2{##1 (##2)}\@internalcite{fpc:81}. In Fig.~\ref{figBC} (panels a-c) we show the comparison between the empirical $BC_{V}$ values derived for three template clusters (M92, NGC3201, and 47~Tuc) spanning almost the whole metallicity range of galactic GC and the $BC_{V}$ scale from our theoretical isochrones. After correcting for the slightly different (by 0.04 mag) zero points (i.e.\ to identical solar bolometric correction), we find very good agreement between the two sets of $BC_{V}$. Also the $BC_{V}$ scales derived from the ATLAS9, Next-Gen (for the main sequence) and LCB transformations agree well with the BK one. In particular, once the $BC_{V}$ zero point is calibrated as we did for BK (see above and SW98), the different scales agree within 0.05 mag all along the isochrones and on the ZAHB; moreover, the brightness differences between TO and ZAHB for fixed age and metallicity agree again within 0.05 mag. This confirms once more the consistency of the absolute ages derived in SW97 and SW98. As a final demonstration of the usefulness of the BK-transformations applied to our isochrones, we show in Figs.~\ref{figM68bv} and \ref{fig47Tucbv} the fits to M68 and 47~Tuc in the $V$-$(B-V)$ plane with our extended isochrones (in SW97 and SW98 only fits up to the HB-brightness were displayed). The data for M68 (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{wal:94}) are the same as in our previous papers, while we used the latest photometry by \def\astroncite##1##2{##1 (##2)}\@internalcite{kws:98} for 47~Tuc. We briefly recall that the ages of the clusters are derived from the brightness difference between the ZAHB (defined as the lower envelope of the observed HB stellar distribution; see SW97 for more details) and TO; this permits to derive the cluster age independently of the knowledge of the cluster distance and reddening (due to the horizontal nature of the HB), but of course the theoretical ZAHB models provide a distance. Therefore, the distance modulus is derived by comparing the theoretical ZAHB level to the observed one, while the reddening is obtained from fitting the unevolved theoretical MS to the cluster one. For 47~Tuc we recover with the new photometric data exactly the same age, reddening and distance modulus as in SW98, where we were using the older composite photometry by \def\astroncite##1##2{##1 (##2)}\@internalcite{hhv:87}. Note that we have used a rather high helium content of $Y=0.273$ for 47~Tuc; this is, at least partially, based on the requirement of consistent vertical and horizontal (colour-dependent) age indicators (see SW98 for a full discussion on this matter). It is evident from the pictures that the extended isochrones reproduce well the observed RGB of both clusters. \begin{figure} \centerline{\includegraphics[draft=false,scale=0.50]{figM68bv.ps}} \caption[]{The $V-(B-V)$--CMD of M68 (data from \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{wal:94}) and isochrones of 10--12 Gyr. For stars fainter than 19th mag only the ridge line is shown; for stars brighter than that only the data points} \protect\label{figM68bv} \end{figure} \begin{figure} \centerline{\includegraphics[draft=false,scale=0.50]{fig47Tucbv.ps}} \caption[]{The $V$-$(B-V)$-CMD of 47~Tuc (data from \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{kws:98}) with isochrones of 8--10 Gyr. For MS stars only the ridge line is shown.} \protect\label{fig47Tucbv} \end{figure} \subsection{Isochrones in the $V-(V-I)$ CMD} For transforming our isochrones to the $V-(V-I)$ plane we have tested all transformations mentioned above. The BK transformations to $(V-I)$ colours do not extend to temperatures higher than 6000 K such that neither ZAHB nor TO models are covered for all metallicities and ages. Moreover, it is not guaranteed automatically that if this set of models reproduces well the observations in one colour band, the same is true for any other colour. \begin{figure}[ht] \centerline{\includegraphics[draft=false,scale=0.60]{figMSvi.eps}} \caption[]{Upper panel: Result of various colour-transformations (indicated in the figure) applied to a theoretical isochrone of ${\rm [Fe/H]} = -1.6$ and 10~Gyr; shown is the main-sequence part. The various labels are explained in Sect.~2.1. Lower panel: The same comparison, but now after adding a constant to the various relations such that they reproduce $(V-I)_\odot=0.73$} \protect\label{figMSvi} \end{figure} To exemplify the comparison between the different $T_{\rm eff}-(V-I)$ relations, we have used a representative isochrone (${\rm [Fe/H]} = -1.6$/10~Gyr), transformed to the $M_V$--$(V-I)$ plane in Fig.~\ref{figMSvi} (upper panel) for the MS and in Fig.~\ref{figRGBvi} for the RGB. In the case of the BCM98 empirical transformations, we have used the empirical $T_{\rm eff}-(V-K)$ relation they give (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{rjw:80}) together with their empirically determined $(V-I)-(V-K)$ relations. As they discuss in that paper, the \def\astroncite##1##2{##1 (##2)}\@internalcite{rjw:80} relation is obtained for solar metallicity stars, but there are good indications that it is not affected by the metallicity. As for M98, the authors derive an empirical $T_{\rm eff}-(V-K)$ relation for metal-rich stars; following the arguments in BCM98, it appears to be safe to use the same relation for the most metal-poor stars, too. In conjunction with the empirical $(V-I)-(V-K)$ relation we obtain again a direct $T_{\rm eff}-(V-I)$ transformation which is displayed in Fig.~\ref{figRGBvi}. \begin{figure} \centerline{\includegraphics[draft=false,scale=0.50]{figRGBvi.eps}} \caption[]{As the upper panel of Fig.~\ref{figMSvi}, but for the red giant branch and ZAHB} \protect\label{figRGBvi} \end{figure} It immediately is evident from these illustrations that colour differences up to between 0.05 and 0.10 mag exist all along the isochrone and that there are no two transformations yielding the same result for all evolutionary stages. This problem persists for other metallicities as well and is, of course, not restricted to our isochrones, but would be similar for any other. Part of the mismatch between the different transformations might be attributed to differences in the composition (of model atmospheres or observed stellar samples), to different or inappropriate assumptions in theory (e.g.\ convection theory, line lists, LTE-conditions) and to the internal spread in the empirical relations. Since the width of the upper MS and the RGB in high-quality photometry is at most as large as 0.05 mag, it is evident that whatever transformation is chosen as it was published, isochrones can deviate from the CMD structures by at least the width of the branches in colour. {\em It is therefore an illusion to assume that a perfect isochrone fitting is a strong indication for excellent theoretical stellar models or isochrones!} Recently, \def\astroncite##1##2{##1 (##2)}\@internalcite{bcp:98} have independently compared a number of theoretical colour transformations with empirical relations for MS stars between spectral types O and M. Since they were interested in aspects of population synthesis, they concluded that there is in general an excellent agreement between theoretical and empirical $T_{\rm eff}$--colour, bolometric correction--colour, and colour--colour relations. However, inspecting their figures (e.g.\ Fig.~6, where the $T_{\rm eff}$--$(V-I)$ relation for the lower MS is shown) one realizes that the deviations between the different sources are of same order as in our case. The same spread is also visible in the empirical $T_{\rm eff}$--$(V-I)$ relation (Fig.~2 as corrected in their Erratum). Only due to the enormous scale of the colour axis the agreement appears to be excellent. Globally, over all temperatures, the agreement is indeed very good, but not sufficient for such narrow structures as the RGB in galactic GC. Along the MS, the empirical relation by \def\astroncite##1##2{##1 (##2)}\@internalcite{aam:96} does not support any theoretical transformation. The Next-Gen and ATLAS9 (BCP98) colours are the most similar, the average differences being within 0.02 mag. Their $T_{\rm eff}-(B-V)$ relations, however, are almost identical in the range of $T_{\rm eff}$ we are dealing with, and in very good agreement with the $T_{\rm eff}-(B-V)$ relation adopted in SW97 and SW98, at least for $T_{\rm eff}$ larger than $\approx 5300$~K. Assuming $(V-I)_\odot = 0.73$ as in \def\astroncite##1##2{##1 (##2)}\@internalcite{bcp:98} we find that this value is reproduced within 0.01 mag by BCP98, while Next-Gen, BK92, AAM96 and LCB yield colours bluer by, respectively, $\approx$ 0.02, 0.02, 0.05, and 0.03 mag. Within the uncertainty of solar colours (see, e.g. the discussions in \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{chmi:81} and \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{tayl:94}) all these values are probably acceptable; nevertheless, for gaining more insight about the differences among the $T_{\rm eff}-(V-I)$ relations under discussion, we normalized the different colour transformations to the same value $(V-I)_\odot = 0.73$, by correcting for the mentioned differences. After performing the necessary shifts one obtains the lower panel of Fig.~\ref{figMSvi} (we have also tested the case with ${\rm [Fe/H]} = -0.7$, with similar results). BCP98, Next-Gen (theoretical relations) and AAM96 (an empirical one) now agree rather well with each other along the whole MS, while the other colours differ from these by different amounts at different metallicities. This latter fact implies that the derivative $\delta (V-I)$/$\delta [Fe/H]$ at a fixed value of the MS brightness, a fundamental quantity needed for applying the MS--fitting method in the $V$-$(V-I)$ plane, does depend on the particular colour transformation used. From the results of these comparisons for the MS phase (which are independent of the underlying theoretical isochrones) it results that BCP98, Next-Gen and AAM96 $T_{\rm eff}-(V-I)$ relations reproduce acceptably well the solar constraint and, moreover, show the same differential behaviour. If we base our 'objective' choice of the colour transformations on the consistency with empirical constraints and with independent theoretical determinations, we have to conclude that BCP98, Next-Gen or AAM96 fulfill these criteria for the MS. While on the MS the isochrones transformed using different $T_{\rm eff}-(V-I)$ relations run more or less parallel to each other (except for LCB), along the RGB they differ also in slopes. However, we find that the two empirical relations tested on the RGB (M98 and BCM98) agree with each other rather well. \def\astroncite##1##2{##1 (##2)}\@internalcite{bcm:98} used the $T_{\rm eff}$--$(V-K)$ relation by \def\astroncite##1##2{##1 (##2)}\@internalcite{rjw:80} and their empirical $(V-I)$--$(V-K)$ relation. These we combined to obtain the $T_{\rm eff}$--$(V-I)$ transformation. In Fig.~\ref{figRGBemp} we show how this relation (``BCM98-Ridgway'') compares to one using the more recent \def\astroncite##1##2{##1}\@internalcite{ben:93} (\def\astroncite##1##2{##2}\@internalcite{ben:93}; ``DB93'') $T_{\rm eff}$--$(V-K)$ relation instead and to the empirical one by \def\astroncite##1##2{##1 (##2)}\@internalcite{mfo:98}. The agreement is, except for the brightest RGB part, better than 0.02 mag. The very good reciprocal consistency between these two purely empirical and independently determined relations is appealing, and on this ground we decided to try to use them for our isochrones. In particular, we selected the BCM98-Ridgway relation which contains an explicit (albeit very weak) metallicity dependence. \begin{figure}[ht] \centerline{\includegraphics[draft=false,scale=0.45]{figRGBemp.ps}} \caption[]{Comparison of RGB empirical $T_{\rm eff}-(V-I)$ relations applied to our isochrones (see text)} \protect\label{figRGBemp} \end{figure} \begin{figure} \centerline{\includegraphics[draft=false,scale=0.45]{figZAHB.eps}} \caption[]{The ZAHB at two metallicities in $(V-I)$: shown is our own relation (``empirical''; see text) and two other transformations. Except for the most extreme colours the agreement is good} \protect\label{figZAHB} \end{figure} We finally need a transformation for our HB models. We have derived an empirical $(B-V)-(V-I)$ relation using multicolor photometries of clusters spanning the relevant range of metallicities. We used HB data for M68 (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{wal:94}), M3 (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{fcc:97}), M5 (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sbs:96}), and 47~Tuc (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{kws:98}), and we applied this relation to our ZAHB models in the $(B-V)$ plane. In Fig.~\ref{figZAHB} the comparison between the ZAHB transformed by using this empirical colour-colour relation and two theoretical transformations is shown for two different metallicities. Evidently, the theoretical relations agree quite well with each other in predicting the turn-down of the ZAHB in the blue. But there are differences not visible due to the horizontal nature of the HB. The predicted colour along the HB differs in some cases by 0.05--0.10 mag, just as for the RGB. Also, the ATLAS9 transformation yields consistently redder stars at the cool edge (similar to the known effect on the RGB) with increasing metallicity. The empirical relation deviates at the blue end, possibly because of the effect that evolved stars show up in the relation. Since observed HBs usually have a large scatter at the blue, less luminous end, this part is not putting any strong constraint on the ZAHB level, such that the difference between the transformations as demonstrated in Fig.~\ref{figZAHB} is not really significant. Except for the mentioned problem of too red colours from the ATLAS9 transformation, we consider all relations as being effectively equivalent. Since we have used empirical data from those clusters we will discuss below, we will use the empirical relation throughout the remainder of this paper. \section{Globular cluster test cases} In the last section we presented a possible choice of colour transformations for all CMD-branches. This choice was made on the basis of which relations agree best with each other. The real justification of this choice, however, will be whether our transformation successfully fulfills the following requirements we are imposing: (i) for a given cluster, the fits in $(B-V)$ and $(V-I)$ should be of the same quality; (ii) the reddening obtained from both colours should be consistent; (iii) derived quantities as age and distance should not depend on the colour used. This implies that the final choice of transformations depends on the set of isochrones used. In the following tests we have employed the BCP98 colours for the MS; between MS and RGB we smoothly switch from the BCP98 $T_{\rm eff}-(V-I)$ relation to the BCM98-Ridgway one, which overlap over a sufficiently wide range. There is no need for any additional correction or shift. For lowest metallicities they almost agree (at the level of $\approx$ 0.01 mag). To test our $T_{\rm eff}-(V-I)$-relation we use four GC of different metallicity for which we have determined the age already in our previous papers and for which $BVI$ data exist. We transform the corresponding isochrones into $(V-I)$ by our new relation and determine the reddening $E(V-I)$. For the transformation of reddenings we take the extinction law by \def\astroncite##1##2{##1 (##2)}\@internalcite{ccm:89}, which gives $E(V-I) = 1.3 \, E(B-V)$. As an example for metal-poor clusters we use M68 (${\rm [Fe/H]} = -1.99\pm0.10$; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{cg:97}); the $(V-I)$ data are from \def\astroncite##1##2{##1 (##2)}\@internalcite{wal:94}. The $(B-V)$-CMD has been shown in Fig.~\ref{figM68bv} and the corresponding $(V-I)$-CMD is displayed in Fig.~\ref{figM68vi} (for sake of clarity the ridge-line only is displayed for the MS). The age determined in SW98 from the $V$ brightness difference between TO and ZAHB is $11.4\pm1.0$ Gyr, in good agreement with the 11 Gyr isochrone displayed Fig.~\ref{figM68vi}. The MS ridge-line was determined by us from the original data by taking the mode of the colour-distribution within brightness bins. The transformed isochrones appear to become too blue in the brightest RGB parts, while the corresponding part in $(B-V)$ is fitting very well\footnote{Worthey (1998, private communication) points out that the \def\astroncite##1##2{##1 (##2)}\@internalcite{wal:94} RGB-data are outliers in $(V-I)$ vs.\ $(V-K)$ when compared to the standard giant branches of \def\astroncite##1##2{##1 (##2)}\@internalcite{dca:90} and the BCM98 data}. The reddening (obtained, as described in \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sw:98}) is 0.06 mag in $(B-V)$ and 0.08 mag in $(V-I)$; their ratio is as predicted by \def\astroncite##1##2{##1 (##2)}\@internalcite{ccm:89}. \begin{figure} \centerline{\includegraphics[draft=false,scale=0.45]{figM68vi.ps}} \caption[]{The $V-(V-I)$ isochrones for M68. Data are from \def\astroncite##1##2{##1 (##2)}\@internalcite{wal:94} and isochrone parameters as in \def\astroncite##1##2{##1 (##2)}\@internalcite{sw:98}. For simplicity, only data points brighter than $V=18.5$ are shown. Large filled circles mark the ridge-line} \protect\label{figM68vi} \end{figure} As an example of an intermediate metal-poor cluster, Fig.~\ref{figM3} shows the comparison between the ridge-line from \def\astroncite##1##2{##1 (##2)}\@internalcite{jb:98} and the 10 Gyr-isochrone for M3 (${\rm [Fe/H]} = -1.34 \pm 0.06$; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{cg:97}), for which SW98 determined differentially the age from $(B-V)$-data (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{fcc:97}) to be $10.1\pm 1.1$ Gyr. The agreement is excellent except for a small deviation in the upper part of the RGB. The derived $E(V-I)$-reddening is 0.03 mag, in perfect agreement with $E(B-V)=0.02$ (SW98; see also \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{bcb:94}). Using the ATLAS9 $T_{\rm eff}-(V-I)$ relation for the RGB as well results in a discrepancy of $\approx$0.05 mag for the larger part of the upper part of the CMD. This difference, if taken seriously, could be taken as evidence that the cluster should be older. \begin{figure} \centerline{\includegraphics[draft=false,scale=0.45]{figM3.eps}} \caption[]{The $V$--$(V-I)$ 10-Gyr isochrone and the ridge line for M3. Data are from \def\astroncite##1##2{##1 (##2)}\@internalcite{jb:98}. Also displayed (dashed) is the same isochrone transformed by using the ATLAS9 colours for the RGB as well} \protect\label{figM3} \end{figure} \begin{figure} \centerline{\includegraphics[draft=false,scale=0.50]{figM5vi.ps}} \caption[]{The $V$--$(V-I)$ ridge-line for M5 (from \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sbh:98}) and isochrones of 8, 9 and 10 Gyr; upper panel: ${\rm [Fe/H]}=-1.0$; lower: ${\rm [Fe/H]}=-1.3$} \protect\label{figM5vi} \end{figure} From the third of our metallicity groups (see SW97) we have selected M5; the absolute cluster age as derived from the $V-(B-V)$ diagram (\@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sbs:96}) is $9.9\pm 0.7$ Gyr (SW98). Very recently \def\astroncite##1##2{##1 (##2)}\@internalcite{sbh:98} have published a new $V-(V-I)$ diagram of M5, and found some discrepancy with respect to the scale and zero points of the \def\astroncite##1##2{##1 (##2)}\@internalcite{sbs:96} $V$ and $I$ magnitudes. However, their estimated TO-luminosity agrees with the one given by \def\astroncite##1##2{##1 (##2)}\@internalcite{sbs:96} within 0.01 mag, and the same is true for the $V$ brightness of the HB and RGB stars in common (with $V$ ranging between 14.5 and 16.5 mag). Therefore, the absolute age as determined from the $V$ difference between ZAHB and TO remains basically unchanged, and the same is true for the cluster distance modulus determined from the fit of theoretical ZAHB sequences to the observed one (in the following we will therefore use the observed ZAHB-$V$ brightness as determined by SW98 employing the more populated HB in the diagram by \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{sbs:96}). The \def\astroncite##1##2{##1 (##2)}\@internalcite{sbh:98} $V-(V-I)$ diagram (they provide the ridge-line) is shown in Fig.~\ref{figM5vi}. The two panels contain the comparison with two of our isochrones, whose metallicities bracket the cluster metallicity ${\rm [Fe/H]} = -1.11 \pm 0.11$, as determined by \def\astroncite##1##2{##1 (##2)}\@internalcite{cg:97}; the distance modulus is fixed by the observed ZAHB brightness. The quality of the fit is good in both cases; there is only a difference in the upper part of the RGB, where the more metal-rich isochrones are too red (as expected). By interpolating for the value of the cluster metallicity we get $E(V-I)=0.03$, in good agreement with $E(B-V)=0.02$ as determined by SW98 (due to a misprint in Table~2 of SW98, the apparent cluster distance modulus was quoted $(m-M_{V})=14.53$, while the correct value is $(m-M_{V})=14.57$). \begin{figure} \centerline{\includegraphics[draft=false,scale=0.50]{fig47Tucvi.ps}} \caption[]{The $V-(V-I)$-CMD of 47~Tuc (data from \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{kws:98}) with the same isochrones as in Fig.~\ref{fig47Tucbv}} \protect\label{fig47Tucvi} \end{figure} Finally, we compare the data by \def\astroncite##1##2{##1 (##2)}\@internalcite{kws:98} for 47~Tuc (${\rm [Fe/H]} = -0.70\pm0.07$; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{cg:97}) with our isochrones. Since in SW98 we used the data by \def\astroncite##1##2{##1 (##2)}\@internalcite{hhv:87} for determining the age from the $(B-V)$ data, we showed the same fit for the new data in Fig.~\ref{fig47Tucbv} (the ridge-line shown for the cluster MS is derived as previously described). The theoretical isochrones are identical to those in SW98 (including the higher helium content of $Y=0.273$) except for the fact that they are now extended to the tip of the RGB. The newly derived age, distance and reddening are the same as in SW98. In Fig.~\ref{fig47Tucvi} the corresponding $(V-I)$-diagram is shown. The reddening determined in this colour is 0.08, which, when transformed to $E(B-V)$ is 0.06, differing from the expected value of 0.05 by only 0.01 mag. Note that while in $(V-I)$ the isochrone appears to be a bit too blue for the lower RGB, it is too red by the same amount in $(B-V)$. The fit is of the same quality in both colours and slightly better than that shown in SW98 for the \def\astroncite##1##2{##1 (##2)}\@internalcite{hhv:87} data. Inspecting the HB of 47~Tuc in $(V-I)$, one has the impression that is inclined towards the blue. This effect is not so evident in $(B-V)$ (Fig.~\ref{fig47Tucbv}). We think it results from the fact that not all stars have both $(B-V)$ and $(V-I)$ colours and that in $(V-I)$ a number of redder HB stars are missing, thus making the inclination, which we ascribe to evolution on the HB, more apparent. In any case, the inclination could only be a brightness effect, independent of colour. To conclude this section, we have demonstrated that the $T_{\rm eff}$--$(V-I)$ transformation we have constructed on the basis of two existing transformations, if applied to our own isochrones (SW98), results in cluster-CMD fits in $(V-I)$ which are equally good as in $(B-V)$ for all metallicities and yield consistent reddenings in the two colours. Thus our requirements formulated at the end of the introduction and the beginning of this section are fulfilled. \section{Conclusions} In this paper we tried to approach the problem of fitting theoretical isochrones to globular cluster data in $(V-I)$. This is important since more and more $I$-band photometric data are becoming available and for some objects (e.g.\ bulge clusters) will be the only data of quality possible. Absolute ages of clusters determined by the turn-off brightness or the $\triangle V$-method (see SW98) depend almost exclusively on the bolometric correction $BC_{V}$. Theoretical values agree, when properly calibrated, with empirical values and also between different transformation sources. On the other hand, transformations of effective temperatures to colours differ from source to source by a rather constant level of 0.05--0.10 mag. From the comparisons we have shown it is evident that (i) there is no straightforward way to decide which transformation is the correct or best one and (ii) that the accuracy with which an isochrone fits a colour-magnitude-diagram does not indicate the quality of the underlying stellar evolution calculations, if the mismatch is of the order of 0.05--0.1 mag. While in Sect.~2 we displayed $(V-I)$--fits, the situation is the same in $(B-V)$ and corresponding examples and conclusions can be found in the literature (e.g.\ \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{gva:98}; \@ifstar{\def\astroncite##1##2{##2}\@internalcite}{\def\astroncite##1##2{##1 ##2}\@internalcite}{dcm:97}). The problem of inaccurate colours is not just a cosmetic one, but can influence absolute age determinations which make use of colour differences (e.g.\ that between turn-off and red giant branch), distances derived by means of the main-sequence--fitting technique, or integrated colours of unresolved stellar populations. Finally, {\em if} a set of transformation was available which is known to reproduce colours accurately, any deviation of an isochrone from the observed CMD will indicate a real inconsistency in the isochrone. As an example we repeat that a colour mismatch of the RGB indicates (but does not prove) a higher helium content for 47~Tuc (SW98). Thus, reliable colour transformations are in principle a source for obtaining additional knowledge about cluster stars. We therefore attempted to find a preliminary solution to this problem, such that we can continue our work about cluster dating with $(V-I)$ data, expecting improved theoretical transformations in the meantime. We found that for MS stars the BCP98, Next-Gen and AAM96 $T_{\rm eff}-(V-I)$ relations reproduce acceptably well the solar constraint and, moreover, show the same differential behaviour with a rather constant offset of a few hundredths of a magnitude. We selected the BCP98 colours for our models. On the red giant branch, no theoretical transformation is in agreement with empirical data over all the metallicity range spanned by our isochrones, but two recent empirical sources (M98 and BCM98) confirm each other quite well. Since BCM98 provide metallicity-dependent relations, we use this one for the RGB. Both parts (MS and RGB) can be connected smoothly along the subgiant branch. For the ZAHB colours we have derived an empirical $(B-V)-(V-I)$ relation using multicolor photometries of clusters spanning the relevant range of metallicities, and we applied this relation to our ZAHB models in the $(B-V)$ plane. While the arguments just given have only led the way to our rather pragmatic combination of two $T_{\rm eff}$--$(V-I)$ transformations, the justification for the combined and final relation comes from the tests we have performed using our own isochrones. For each of the four metallicity ranges defined in SW97, which reach from the most metal-poor halo to the half-solar metallicity disk clusters, we selected one cluster with $BVI$-photometry (three of them had already been used for absolute age determinations in SW98). We then showed successfully that (i) the quality of the isochrone fit is comparable in both colours and (ii) the reddenings determined independently from these two fits fulfill the reddening law by \def\astroncite##1##2{##1 (##2)}\@internalcite{ccm:89} either exactly or within 0.01 mag. We therefore conclude that the combination of the (theoretical) transformation of BCP98 for the main-sequence and the (empirical) one by BCM98 for giants is of sufficient accuracy as to allow us isochrone fits to globular cluster colour-magnitude-diagrams in $V$--$(V-I)$. In a forthcoming paper we will determine cluster ages based on $VI$-photometry, both for clusters we already have investigated in our previous papers and for clusters with $VI$-data only. Although $(V-I)$ is less sensitive to metallicity than $(B-V)$, we found that $\triangle (V-I)$ (the colour difference between turn-off and red giant branch, which we need for relative age determinations) at fixed age is actually {\em more} metal-sensitive. Our approach of grouping clusters into four metallicity bins will therefore not be accurate enough. Either a finer metallicity grouping is needed (implying a higher number of clusters for absolute age determination), or a purely relative age approach is needed as in \def\astroncite##1##2{##1 (##2)}\@internalcite{psw:98}. Independent of this, we are still waiting for improved theoretical colour transformations, because for more metal-rich $\alpha$-enhanced clusters (e.g.\ bulge clusters) the BCM98 results cannot be applied straightforwardly. The study of such clusters therefore actually requires to some extent an extrapolation of our adopted colour transformations. \begin{acknowledgements} We are grateful to L.~Girardi and I.~Steele for helpful discussions and E.~M\"uller for a careful reading of the manuscript. We thank T.~Lejeune for providing an updated version of his transformations in electronic form, and an anonymous referee for his comments, which helped to improve this paper. \end{acknowledgements} \clearpage
\section{Introduction} Double beta decay has been long recognized as a powerful tool for the study of lepton conservation. The theoretical description of $\beta\beta$ decay involves particle physics and nuclear structure physics. The latter is the topic of the present work. After a brief introduction and a review of the experimental situation I will describe three distinct set of problems: \begin{itemize} \item $2\nu$ decay: the physics of the Gamow-Teller amplitudes. \item $0\nu$ decay - exchange of light massive Majorana neutrinos: no selection rules on multipoles, role of nucleon correlations, sensitivity to nuclear models. \item $0\nu$ decay - exchange of heavy neutrinos: physics of the nucleon-nucleon states at short distances. \end{itemize} Since the lifetimes of $\beta\beta$ decay are so long, the experimental search has spawned a whole field of experiments requiring very low background. During the last decade there has been enormous progress in the experimental study of the double beta decay. The $2\nu$ decay, with lifetimes of $10^{19} - 10^{21}$ y, has been now observed in ten cases, often repeatedly and by different groups \cite{Morales}. The halflives are typically determined to better than 10\% accuracy. Thus, the $2\nu$ decay has become a tool, rather than an exotic curiosity. The main experimental effort is concentrated, naturally, on the search for the $0\nu$ decay. There, multikilogram sources with very low background are the state of the art at present. Various techniques are being pursued: two large experimental efforts (Heidelberg-Moscow collaboration in Gran Sasso and the IGEX collaboration in Canfranc) employ enriched $^{76}$Ge in sets of detectors operated deep underground. The latest result \cite{Baudis} is based on 41.55 kg y of exposure and halflife limit $T_{1/2} > 1.3 \times 10^{25}$y (90\% CL). Alternatively, the analysis of the 24.16 kg y of data with pulse shape measurement, and hence improved background suppression, leads to even better limit of $T_{1/2} > 1.6 \times 10^{25}$y (90\% CL). Note, that in the latter case the actual number of observed events is smaller than the expected background. When taking advantage of this, the authors of \cite{Baudis} report yet much improved limit of $T_{1/2} > 5.7 \times 10^{25}$y (90\% CL). The latest IGEX limit is $T_{1/2} \ge 0.8 \times 10^{25}$y (90\% CL)\cite{IGEX}. I will discuss the interpretation of this result in terms of the neutrino Majorana mass below in Section 3. Other techniques which allow expansions to multikilogram sources involve gas TPC ($^{136}$Xe in the Gotthardt tunnel \cite{Xe}), electron tracking detectors combined with calorimeters (NEMO \cite{Nemo}, ELEGANTS \cite{Ejiri}), and cryogenic bolometers \cite{Fiorini}. These experiments will make it possible (or currently already achieve) to reach neutrino mass limit well below 1 eV. In addition, there are plans for improvements of the limit to the 0.1 eV range by scaling up the source mass to about one ton quantities. Such experiments, with either enriched $^{76}$Ge (GENIUS) or with large amount of cooled TeO$_2$ (CUORE) will require very large capital expenditures and running times of about five years. The field of $\beta\beta$ decay then becomes competitive, in complexity, manpower, and price, to large accelerator particle physics experiments. The Majorana neutrino mass $\langle m_{\nu} \rangle$ below 1 eV is an important landmark. This has been stressed often, e.g., in the work of Georgi and Glashow \cite{GG}. There, the authors quote as ``established facts'' the discovery by the SuperKamiokande collaboration of the $\nu_{\mu} \rightarrow \nu_{\tau}$ neutrino oscillations with $\Delta m^2 \simeq 10^{-3}$ eV$^2$ and mixing angle $\sin^2 2 \theta \simeq 1$, and the solar neutrino deficit, confirmed in five experiments, which involves $\nu_e \rightarrow \nu_x$ oscillations with $\Delta m^2 \le 10^{-5}$ eV$^2$. The ``tentative facts'' are based on theoretical prejudices and need experimental confirmation. They are the existence of precisely three massive Majorana neutrinos, and the assumption that these neutrinos are responsible for the hot dark matter, leading to the conclusion that $m_1 + m_2 + m_3 \sim 6 $ eV. Taking all of that together one is forced to the wholly unexpected and somewhat bizarre conclusion that the three neutrino flavors are essentially degenerate with masses of 2 eV each (albeit with a sizable uncertainty in this value). It is now clear that if the study of the $0\nu$ $\beta\beta$ decay can without doubt establish that $\langle m_{\nu} \rangle < 1$ eV, this finding has a profound consequences for the structure of the neutrino mixing matrix. In particular, as Georgi and Glashow \cite{GG} argue, it would lead to maximum mixing involving electron neutrinos also. \section{Two neutrino decay} This decay, characterized by the transformation of two neutrons into two protons with the emission of two electrons and two $\bar{\nu}_e$, does not violate any selection rules. Since the energies involved are modest, the allowed approximation should be applicable, and the rate is governed by the double Gamow-Teller matrix element \begin{equation} M_{GT}^{2\nu} = \sum_m \frac{\langle f || \sigma \tau_+ || m \rangle \times \langle m || \sigma \tau_+ || i \rangle } { E_m - (M_i + M_f)/2 } ~' \end{equation} where $i, f$ are the ground states in the initial and final nuclei, and $m$ are the intermediate $1^+$ (virtual) states in the odd-odd nucleus. The first factor in the numerator above represents the $\beta^+$ (or ($n,p$)) amplitude for the final nucleus, while the second one represents the $\beta^-$ (or ($p,n$)) amplitude for the initial nucleus. Thus, in order to correctly evaluate the $2\nu$ decay rate, we have to know, at least in principle, {\it all} GT amplitudes for both $\beta^-$ and $\beta^+$ processes, including their signs. The difficulty is that the $2\nu$ matrix element exhausts a very small fraction ($10^{-5} - 10^{-7}$) of the double GT sum rule \cite{double}, and hence it is sensitive to details of nuclear structure. \begin{figure}[h] \begin{center} \mbox{\psfig{figure=ca48.ps,width=5cm}} \caption{The $\beta^-$ strength (upper panel), and the contributions to the $2\nu$ matrix element, eq. (1) (lower panel). Both as the function of the excitation energy in $^{48}$Sc.} \label{fig:ca48} \end{center} \end{figure} Various approaches used in the evaluation of the $2\nu$ decay rate have been reviewed recently by Suhonen and Civitarese \cite{SC}. The Quasiparticle Random Phase Approximation (QRPA) has been the most popular theoretical tool in the recent past. Its main ingredients, the repulsive particle-hole spin-isospin interaction, and the attractive particle-particle interaction, clearly play a decisive role in the concentration of the $\beta^-$ strength in the giant GT resonance, and the relative suppression of the $\beta^+$ strength and its concentration at low excitation energies. Together, these two ingredients are able to explain the suppression of the $2\nu$ matrix element when expressed in terms of the corresponding sum rule. Yet, the QRPA is often criticized. Two ``undesirable'', and to some extent unrelated, features are usually quoted. One is the extreme sensitivity of the decay rate to the strength of the particle-particle force (often denoted as $g_{pp}$). This decreases the predictive power of the method. The other one is the fact that for a realistic value of $g_{pp}$ the QRPA solutions are close to their critical value (so-called collapse). This indicates a phase transition, i.e., a rearrangement of the nuclear ground state. QRPA is meant to describe small deviations from the unperturbed ground state, and thus is not fully applicable near the point of collapse. Numerous approaches have been made to extend the range of validity of QRPA, see, e.g., Ref. \cite{SC}. The description of all these generalizations is beyond the scope of this talk. \begin{table}[h] \caption{ Experimental halflives of the $2\nu$ decay, and the ratios of calculated to experimental halflives (see text).} \begin{center} \begin{tabular}{lllc} \hline\hline & $\frac{T_{1/2}^{calc}}{T_{1/2}^{exp}}$ & $\frac{T_{1/2}^{calc}}{T_{1/2}^{exp}}$ & $T_{1/2}$ (y) \\ Nucleus & QRPA &shell model & exp. \\ \hline $^{48}$Ca& -- & 0.91 & 4.3$\times 10^{19}$ \\ $^{76}$Ge& 0.71 & 1.44 &1.8$\times 10^{21}$ \\ $^{82}$Se& 1.5 & 0.46 &8.0$\times 10^{19}$ \\ $^{100}$Mo& 0.6 & -- &1.0$\times 10^{19}$ \\ $^{128}$Te& 0.27 & 0.25 &2.0$\times 10^{24}$ \\ $^{130}$Te& 0.27 & 0.29 &8.0$\times 10^{20}$ \\ $^{136}$Xe& $<$ 1.5 & $<$ 3.7 &$>$5.6$\times 10^{20}$ \\ \hline \label{tab:2nu} \end{tabular} \end{center} \end{table} At the same time, detailed calculations show that the sum over the excited states in Eq.(1) converges quite rapidly \cite{EEV}. In fact, a few low-lying states usually exhaust the whole matrix element. Thus, it is not really necessary to describe all GT amplitudes; it is enough to describe correctly the $\beta^+$ and $\beta^-$ amplitudes of the low-lying states, and include everything else in the overall renormalization (quenching) of the GT strength. The situation is illustrated in Fig. \ref{fig:ca48} modified from Ref. \cite{ca48sm}. Nuclear shell model methods are presently capable of handling much larger configuration spaces than even a few years ago. Thus, for many nuclei the evaluation of the $2\nu$ rates within the 0$\hbar \omega$ shell model space is feasible. (Heavy nuclei with permanent deformation, like $^{150}$Nd and $^{238}$U remain, however, beyond reach.) Using the interacting shell model avoids, naturally, the above difficulties of QRPA. At the same time, the shell model is capable to predict, with the same method and the same residual interaction, a wealth of spectroscopic data, allowing a much better test of the predictive power. To judge the degree of understanding of the $2\nu$ decay I show in Table \ref{tab:2nu} the comparison with experiment of the initial Caltech QRPA calculation \cite{Engel} and the modern shell model \cite{Caurier97}. One can see that both methods are able, at least in these cases, explain the $2\nu$ decay rates reasonably well, even though in the case of Te both methods underestimate the halflife by a factor of about four. \section{Neutrinoless decay: light Majorana neutrino} In the neutrinoless decay the two electrons are the only leptons emitted, and consequently their sum energy is just the sharp nuclear mass difference. This feature makes the experimental recognition of the $0\nu$ decay much easier; it also results in a more favorable phase space factor. If one assumes that the $0\nu$ decay is caused by the exchange (virtual) of a light Majorana neutrino between the two nucleons, then several new features arise: a) the exchanged neutrino has a momentum $q \sim 1/r_{nn} \simeq 50 - 100$ MeV ($r_{nn}$ is the distance between the decaying nucleons). Hence, the dependence on the energy in the intermediate state is weak and the closure approximation is applicable. Also, b) since $qR > 1$ ($R$ is the nuclear radius), the expansion in multipoles is not convergent, unlike in the $2\nu$ decay. In fact, all possible multipoles contribute by a comparable amount. Finally, c) the neutrino propagator results in a neutrino potential of a relatively long range. Thus, in order to evaluate the rate of the $0\nu$ decay, we need to evaluate only the matrix element connecting the ground states $0^+$ of the initial and final nuclei. Again, we can use the QRPA or the shell model. Both calculations show that the features enumerated above are indeed present. In addition, the QRPA typically shows less extreme dependence on the particle-particle coupling constant $g_{pp}$, since the contribution of the $1^+$ multipole is relatively small. The calculations also suggest that for quantitatively correct results one has to treat the short range nucleon-nucleon repulsion carefully, despite the long range of the neutrino potential. Does that mean that the calculated matrix elements are insensitive to nuclear structure? An answer to that question has obviously great importance, since unlike the $2\nu$ decay, we cannot directly test whether the calculation is correct or not. For simplicity, let us assume that the $0\nu$ $\beta\beta$ decay is mediated only by the exchange of a light Majorana neutrino. The relevant nuclear matrix element is then the combination $M_{GT}^{0\nu} - M_F^{0\nu}$, where the GT and F operators change two neutrons into two protons, and contain the corresponding operator plus the neutrino potential. One can express these matrix elements either in terms of the proton particle - neutron hole multipoles (i.e. the usual beta decay operators) or in the multipoles coupling of the exchanged pair, $nn$ and $pp$. \begin{figure}[h] \begin{center} \mbox{\epsfig{figure=0nu.eps,width=6cm}} \caption{The cumulative contribution of the pair states with the natural parity multipolarity to the $0\nu$ nuclear matrix element combination $M_{GT}^{0\nu} - M_F^{0\nu}$. The full line is for $^{76}$Ge and the dashed line for $^{48}$Ca.} \label{fig:0nu} \end{center} \end{figure} When using the decomposition in the proton particle - neutron hole multipoles, one finds that all possible multipoles (given the one-nucleon states near the Fermi level) contribute, and the contributions have typically equal signs. Hence, there does not seem to be much cancellation. However, perhaps more physical is the decomposition into the exchanged pair multipoles. There one finds, first of all, that only natural parity multipoles ($\pi = (-1)^I$) contribute noticeably. And there is a rather severe cancellation. The biggest contribution comes from the $0^+$, i.e., the pairing part. All other multipoles, related to higher seniority states, contribute with an opposite sign. The final matrix element is then a difference of the pairing and higher multipole (or broken pair $\equiv$ higher seniority) parts, and is considerably smaller than either of them. This is illustrated in Fig. \ref{fig:0nu} where the cumulative effect is shown, i.e. the quantity \begin{equation} M(I) = \sum_J^I \left[ M_{GT}^{0\nu}(J) - M_F^{0\nu}(J) \right] \end{equation} is displayed for $^{76}$Ge (from \cite{muto2}) and $^{48}$Ca (from \cite{ca48sm}). Thus, the final result depends sensitively on both the correct description of the pairing and on the admixtures of higher seniority configurations in the corresponding initial and final nuclei. Since there is no objective way to judge which calculation is correct, one often uses the spread between the calculated values as a measure of the theoretical uncertainty. This is illustrated in Table \ref{tab:0nu}. There, I have chosen two representative QRPA sets of results, the highly truncated ``classical'' shell model result of Haxton and Stephenson \cite{Haxton}, and the result of more recent shell model calculation which is convergent for the set of single particle states chosen (essentially 0$\hbar \omega$ space). For the most important case of $^{76}$Ge, the calculated rates differ by a factor of 6-7. Since the effective neutrino mass $\langle m_{\nu} \rangle$ is inversely proportional to the square root of the lifetime, the experimental limit of $1.6 \times 10^{25}$ y translates into limits of about 1 eV according to \cite{Engel,Caurier97} and about 0.4 eV according to \cite{Staudt,Haxton}. On the other hand, if one would accept the more stringent limit of $5.7 \times 10^{25}$ \cite{Baudis}, even the more pessimistic matrix elements restrict $\langle m_{\nu} \rangle < 0.5$ eV, hence the scenario discussed by Georgi and Glashow \cite{GG} is confirmed. Needles to say, a more objective measure of the theoretical uncertainty would be highly desirable. \begin{table}[h] \caption{ Halflives in years calculated for $\langle m_{\nu} \rangle$ = 1 eV by various representative methods.} \begin{center} \begin{tabular}{r|rrll} \hline\hline & QRPA \cite{Staudt} & QRPA \cite{Engel} & SM \cite{Haxton} & SM \cite{Caurier97} \\ \hline $^{76}$Ge& 2.3$\times10^{24}$ & 1.4$\times10^{25}$ &2.4$\times10^{24}$ & 1.7$\times10^{25}$ \\ $^{82}$Se& 6.0$\times10^{23}$ & 5.6$\times10^{24}$ & 8.4$\times10^{23}$ & 2.4$\times10^{24}$ \\ $^{100}$Mo & 1.3$\times 10^{24}$ & 1.9$\times 10^{24}$ & -- & -- \\ $^{130}$Te& 4.9$\times10^{23}$ & 6.6$\times10^{23}$ & 2.3$\times10^{23}$ & -- \\ $^{136}$Xe& 2.2$\times10^{24}$ & 3.3$\times10^{24}$ & -- & 1.2$\times10^{25}$ \\ \hline \end{tabular} \end{center} \label{tab:0nu} \end{table} \section{Neutrinoless decay: very heavy Majorana neutrino} The neutrinoless $\beta\beta$ decay can be also mediated by the exchange of a heavy neutrino. The decay rate is then inversely proportional to the square of the effective neutrino mass \cite{Vergados}. In this context it is particularly interesting to consider the left-right symmetric model proposed by Mohapatra \cite{Mohapatra}. In it, one can find a relation between the mass of the heavy neutrino $M_N$ and the mass of the right-handed vector boson $W_R$. Thus, the limit on the $\beta\beta$ rate provides, within that specific model, a stringent lower limit on the mass of $W_R$. The process then involves the emission of the heavy $W_R^-$ by the first neutron, the vertex $W_R^- \rightarrow e^- + \nu_N$ followed by $\nu_N \rightarrow e^- + W_R^+$ with the absorption of the $W_R^+$ on the second neutron. Since all exchanged particles between the two neutrons are very heavy, the corresponding ``neutrino potential'' is of essentially zero range. Hence, when calculating the nuclear matrix element, one has to take into account carefully the short range nucleon-nucleon repulsion. As long as we treat the nucleus as an ensemble of nucleons only, the only way to have a nonvanishing nuclear matrix elements for the above process is to treat the nucleons as finite size particles. In fact, that is the standard way to approach the problem \cite{Vergados}; the nucleon size is described by a dipole form factor with the cut-off parameter $\Lambda \simeq$ 0.85 GeV. However, another way of treating the problem is possible, and already mentioned in \cite{Vergados}. Let us recall how the analogous situation is treated in the description of the parity-violating nucleon-nucleon force \cite{AH}. There, instead of the weak (i.e., very short range) interaction of two nucleons, one assumes that a meson ($\pi, \omega, \rho$) is emitted by one nucleon and absorbed by another one. One of the vertices is the parity-violating one, and the other one is the usual parity-conserving strong one. The corresponding range is then just the meson exchange range, easily treated. The situation is schematically depicted in the left-hand panel of Fig. \ref{fig:moh}. The analogy for $\beta\beta$ decay is shown in the right-hand graph. It involves two pions, and the ``elementary'' lepton number violating $\beta\beta$ decay then involves a transformation of two pions into two electrons. Again, the range is just the pion exchange range. To my knowledge, no detailed evaluation of the corresponding graph was ever made (see, however, Ref.\cite{Savage}). It would be interesting to see if it would lead to a more or less stringent limit on the mass of the $W_R$ than the treatment with form factors. \begin{figure}[h] \begin{center} \mbox{\psfig{figure=mohap.ps,width=6cm}} \caption{The Feynman graph description of the parity-violating nucleon-nucleon force (left graph) and of the $\beta\beta$ decay with the exchange of a heavy neutrino mediated by the pion exchange.} \label{fig:moh} \end{center} \end{figure} \vspace{1.5cm} {\large{\bf Acknowledgment}} It is my pleasant duty to thank Prof. Christianne Miehe for organizing the workshop and inviting me to Strasbourg. The permission of Etienne Caurier and Frederick Nowacki to use their results is greatly appreciated. This work was supported by the US Department of Energy under Grant No. DE-FG03-88ER-40397.
\section{Introduction} \subsection{Background} Compact galactic nuclei are a common but still poorly understood feature of many, if not most spiral galaxies, including our own Milky Way. Such nuclei are regions only a few parsecs or less in diameter, but have complex structures and sometimes produce energy outputs equal to a significant fraction of the total galaxy luminosity. For the present work we define a {\it compact nucleus} as a non-stellar, point-like light enhancement at or near the center of a galaxy whose brightness is in excess of an extrapolation of the galaxy light profile from the surrounding regions (e.g., Phillips {\it et al.}\ 1996) and which is dynamically distinct from the surrounding galaxy disk. Compact nuclei are not simply nuclear \mbox{H\,{\sc ii}}\ regions, although in some cases, they may be one component of a composite nucleus---i.e. they may be embedded in a nuclear \mbox{H\,{\sc ii}}\ region, a nuclear starburst region, or a nuclear disk (e.g., Filippenko 1989; van den Bergh 1995; Ford {\it et al.}\ 1997). Compact galactic nuclei include both ``active galactic nuclei'' or ``AGNs'' (i.e., QSOs, Seyferts, LINERs), which exhibit characteristic emission lines and non-stellar spectral energy distributions, as well as ``compact star cluster nuclei'', whose properties can be accounted for by dense concentrations of stars. Compact star cluster nuclei are the densest known stellar systems, with even modest examples containing a mass of more than 10$^6\mbox{${\cal M}_\odot$}$ within a few parsecs radius (e.g., Lauer {\it et al.}\ 1998; hereafter L98). Star cluster nuclei differ from scaled-up versions of normal dense star clusters in that they exhibit a wider range of stellar ages, quantities of associated dense gas, and sometimes evidence for the the existence of central massive black holes. Possible evolutionary links between AGNs and compact star cluster nuclei are still widely debated (e.g., Norman \& Scoville 1988; Filippenko 1992; Williams \& Perry 1994 and references therein). Compact galactic nuclei can be difficult to pick out even in nearby galaxies. Studies from the ground are limited by the effects of seeing and confusion with bright bulges or regions of enhanced star formation in galaxy centers. Moreover, in the past, due to the limited dynamic range of photographic plates, exposures that revealed the outer structures of galaxies often overexposed the nuclear regions, thus obscuring compact nuclear sources (e.g., van den Bergh 1995). For some time it was accepted that compact nuclei were mainly limited to dwarf spheroidal galaxies (e.g., Binggeli {\it et al.}\ 1984; Binggeli \& Cameron 1991) and to elliptical galaxies or early-type spiral bulges in which massive ($10^{6}-10^{7}$\mbox{${\cal M}_\odot$}) black holes appear to be commonplace (e.g., Ford {\it et al.}\ 1997; van der Marel 1998 and references therein). However, high-quality photographic plates and high dynamic range CCD images, where the nuclear regions of galaxies are not ``burned in'', have helped to change our perceptions on the range in properties of compact nuclei and on the types of galaxies that may harbor them (e.g., van den Bergh 1995). In particular, it has become evident that true compact nuclei are common in low-luminosity, late-type (Scd-Sdm) spiral disks that lack a bulge component (i.e., ``extreme late-type spirals''). Although the compact nuclei of extreme late-type spirals are often relatively faint ($M_{V}\geq\sim -12$), in nearby galaxies they become readily detectable even from the ground under normal ($\sim 1''$) seeing conditions, since confusion from bulges or brilliant circumnuclear disks is minimal, and dust obscuration is often low. For example, Matthews \& Gallagher (1997) noted the existence of moderate-to-low luminosity compact, semi-stellar nuclei in 10 of 49 nearby, extreme late-type spirals they imaged from the ground. In a Snapshot survey with the {\it Hubble Space Telescope} ({\it HST}), Phillips {\it et al.}\ (1996) also found compact nuclei in a number of nearby, late-type, low-luminosity spirals (see also Carollo {\it et al.}\ 1997). Compact nuclei may be common in pure disk galaxies at least to moderate redshifts. Sarajedini {\it et al.}\ (1996) found that in a magnitude-limited survey ($I\leq$21.5), 84 of 825 galaxies contained unresolved nuclear sources. Of these, 57\% of the host galaxies could be adequately modelled by an exponential disk alone, with no bulge component (see also Sarajedini 1996). The bulk of the compact nuclei in nearby extreme late-type spirals appear to be of the ``compact star cluster'' variety (e.g., Shields \& Filippenko 1992; Phillips {\it et al.}\ 1996). Nearby examples include the nucleus of the Local Group Scd spiral M33 (e.g., Kormendy \& McClure 1993; hereafter KM; L98) and the nucleus of the Sd spiral NGC~7793 (D{\'\i}az {\it et al.}\ 1982; Shields \& Filippenko 1992). However, in a ground-based spectroscopic survey of 43 Scd and later nucleated galaxies, Ho (1996) reported that a surprising 16\% of these objects showed evidence for nuclear activity (i.e., Seyfert, LINER, or ``transition'' nuclei). Among the handful of examples of compact nuclei in low-luminosity disk galaxies that have been studied in detail, there have been other surprises. One of the nearest QSOs (0351+026 at $z$=0.036) was found by Bothun {\it et al.}\ (1982a,b) to lie within a faint ($M_{V}=-18.6$),\footnote{$H_{o}$=75~\mbox{km s$^{-1}$}~Mpc$^{-1}$ is assumed throughout this work.} blue, moderately low surface brightness disk galaxy. Moreover, it is interacting with a nucleated, low surface brightness, gas-rich galaxy with $M_{V}\sim-17.5$ (Bothun {\it et al.}\ 1982b). Kunth {\it et al.}\ (1987) were the first to image the galaxy G1200-2038 associated with a Seyfert~2 nucleus. The host galaxy is a very small, faint dwarf galaxy ($D_{25}\approx$6~kpc and $M_{V}=-$16.8). The detailed morphology of this galaxy is uncertain, but it is well represented by a pure exponential disk, and its small physical size and low luminosity are consistent with an extreme late-type spiral galaxy (cf. Matthews \& Gallagher 1997). In spite of the properties of the host galaxy, the nucleus of G1200-2038 has a luminosity $M_{V}\sim-$16.1, yielding a ratio of nuclear to host galaxy luminosity typical of much more luminous AGNs. Such cases as these may hold important clues as to how different ``flavors'' of active nuclei are related, and they raise the question of whether compact nuclei in nearby extreme late-type spirals have special characteristics, or perhaps were much more powerful objects in the past that have since run out of fuel. Furthermore, since moderate-to-low luminosity Sd-Sm spirals are the most common class of disk galaxy (van der Kruit 1987), if low-level nuclear activity was commonplace in these objects, it may make a significant contribution to the soft X-ray background (e.g., Koratkar {\it et al.}\ 1995). However, the majority of nuclei in nearby extreme late-type spirals appear to be either weakly active or non-active, and these galaxies therefore furnish examples of the little-explored low-luminosity regime for galactic nuclei and nuclear activity. The investigation of this faint end of the galactic nuclear activity sequence is critical for constraining the origin and evolution of compact nuclei and nuclear activity, as well as the physical processes that power them, since the existence of compact nuclei in small diffuse galaxies with weak central potentials and no bulge component is difficult to explain in light of current models for the origin of compact galactic nuclei and nuclear activity (see Sect.~6). Nonetheless, exploration of these systems is largely just beginning (e.g., Filippenko \& Sargent 1985; Ho {\it et al.}\ 1995a,1997a,c; Koratkar {\it et al.}\ 1995; Maiolino \& Rieke 1995; Sarajedini {\it et al.}\ 1996). Still unanswered questions include: Do all star cluster nuclei contain central black holes? Are young stellar populations centrally located within nuclei? How do structures and spatial scales compare between active and non-active nuclei? What is the minimum luminosity for an active nucleus? Because confusion with dust, light from the bulge, and bright circumnuclear material are minimized, detailed studies of the nuclei of nearby extreme late-types spirals can help to afford unique insights to many of these problems. By studying the nuclei of extreme late-type spirals, we can also hope to further our understanding of the role of the host galaxy in provoking and sustaining nuclear activity, and explore what relationships may exist between morphology, luminosity, or other global properties of the host galaxy and its nuclear characteristics. For example Filippenko \& Sargent (1989) discovered the faintest Seyfert~1 nucleus yet known ($M_{B}\sim -11$) in NGC~4395, one of the objects we explore further in the present study. We also investigate the `normal' nuclei in the Sd/Sdm galaxies NGC~4242 and ESO~359-029, and analyze the nearby M33 nucleus as a comparison object. \subsection{Past Observational Limitations and Progress from HST} Constraining the physics of compact galactic nuclei requires resolution of their structures and measurements of their physical sizes. However, since half-light radii are typically $\leq$10~pc, even the closest spiral galaxy nuclei are difficult to resolve from the ground with conventional observing techniques (e.g., Gallagher {\it et al.}\ 1982; Nieto {\it et al.}\ 1986; Mould {\it et al.}\ 1989; KM). In general, space-based observations with image quality of $\leq$\as{0}{1} are needed to procure sufficiently detailed information to advance our understanding of these objects. The first such observations were supplied for M31 by the Stratoscope~II balloon-borne telescope (Light {\it et al.}\ 1974). More extensive data on structures of external spiral galaxy nuclei have come from ultraviolet (UV) and optical images obtained with the cameras aboard the {\it HST} that now achieve $\sim$\as{0}{05} angular resolution. Even before its spherical aberration was corrected, the {\it HST}\ displayed its resolving power by revealing that M31 has a double nucleus (Lauer {\it et al.}\ 1993), showing that compact nuclei are common in the UV (e.g., Fabbiano {\it et al.}\ 1994, Maoz {\it et al.}\ 1995), and by providing initial measures of galactic nuclear structures (e.g., Crane {\it et al.}\ 1993; Lauer {\it et al.}\ 1993,1995; King {\it et al.}\ 1995; Ford {\it et al.}\ 1992; Phillips {\it et al.}\ 1996). These results have been confirmed with optical and mid-UV images from the corrected optics of the Wide Field Planetary Camera 2 (WFPC2) and the COSTAR-corrected Faint Object Camera (e.g., Colina {\it et al.}\ 1997, Devereux {\it et al.}\ 1997). Aberration-corrected {\it HST}\ images of the centers of a few nearby spirals show that typical galaxy nuclei are often unresolved in the UV, especially in galaxies with some evidence for activity. In the visible, nuclei of spirals can have complex surroundings (e.g. as in M100), and range in size from $<$1~pc (e.g., M81, as measured by Devereux {\it et al.}\ 1997) to core radii of 1.4~pc and 3.7~pc respectively for the P1 and P2 double nuclei of M31 (Lauer {\it et al.}\ 1993; L98). \subsection{New Goals} In this paper we present a WFPC2 investigation of the galactic nuclei of three extreme late-type, Sd-Sdm spiral galaxies. We also present an analysis of WFPC2 data of the nucleus of our closest Scd spiral neighbor, M33, in order to form a comparative framework for our analysis. All of these galaxies are dominated by their stellar disks and have little or no bulge component. Our objectives are to study the optical structures, colors, and spatial scales of the low-luminosity, ``naked'' galactic nuclei found in these galaxies, and to compare these results with other compact galactic nuclei. A preliminary version of this work was presented by Matthews {\it et al.}\ (1996). The global properties of our program galaxies are described in Table~1. Our first target, NGC~4395, is an SAd~III-IV Seyfert~1 galaxy (Filippenko \& Sargent 1989, Filippenko {\it et al.}\ 1993). NGC~4242 (SABd~III) is morphologically similar to NGC~4395 (cf. Sandage \& Bedke 1994), although slightly more distant, and its optically prominent nucleus does not appear to be active (Ho {\it et al.}\ 1995a). The less luminous, nucleated extreme late-type galaxy, ESO~359-029 (Sdm) was selected from the sample of Matthews \& Gallagher (1997; see also Sandage \& Fomalont 1993) and is one of the lowest luminosity disk galaxies ($M_{B}=-15.1$) known to have a nucleus. ESO~359-029 does not appear to be active, but a longslit spectrum (Matthews 1998) reveals the kinematic signature of a compact central mass concentration at the location of the nucleus. Lastly, M33 (SAcd~III) and its nucleus have similar luminosities to our other targets, but because it is nearby, it is well-resolved and has been previously extensively investigated at multiple wavelengths (see L98). It therefore serves as an excellent comparison object, a check on the validity of our analysis techniques, as well as a guide for the interpretation of our results. \section{Observations} Images of NGC~4395, NGC~4242, and ESO~359-029 were obtained for this program by the WFPC2 Investigation Definition Team. The observations are summarized in Table~2. The nuclei of the galaxies were observed with the Planetary Camera~2 (PC2) which gives a scale of \as{0}{0455} per pixel on an 800$\times$800 pixel Loral CCD. We used a gain of 7~$e^{-}$ per data number (DN), and the CCD was operated at a temperature of $-$88$^{\circ}$~C. More details regarding WFPC2 can be found in e.g., Trauger {\it et al.}\ (1994), Holtzman {\it et al.}\ (1995a), or Biretta {\it et al.}\ (1996). Since our program was limited to one orbit per galaxy, we obtained one short and two moderate, CR-SPLIT exposures for each object in the $F450W$ (WFPC2 broad $B$-band) and $F814W$ (WFPC2 broad $I$-band) filters. The short exposures were made to avoid saturated pixels in the bright centers of the nuclei. Total exposure times for each object are given in Table~2. The $F450W$ filter includes the H$\beta$ and [\mbox{O\,{\sc iii}}] $\lambda\lambda$4959, 5007 emission lines, which can be very strong near an AGN. Together, $F450W$ and $F814W$ give us a good color baseline. The data were reduced following the precepts of Holtzman {\it et al.}\ (1995a) and using the approach discussed by Watson {\it et al.}\ (1996) for removing cosmic rays. Calibrations of magnitudes and fluxes are based on the system described by Holtzman {\it et al.}\ (1995b). The data for M33 were obtained from the archives of the Canadian Astronomical Data Centre. These data consist of recalibrated WFPC2 PC2 images taken in the $F555W$ (WFPC2 $V$-band) and $F814W$ filters, and were pipeline calibrated and combined in IRAF\footnote{IRAF is distributed by the National Optical Astronomy Observatories, which is operated by the Associated Universities for Research in Astronomy, Inc. under cooperative agreement with the National Science Foundation.} to eliminate cosmic rays. The archival observations that we used are also summarized in Table~2. The $F814W$ images of each of our target nuclei are shown in Fig.~1. \section{A Method for Characterizing the Brightness Profiles of the Nuclei} \subsection{Model Fits} In order to obtain information about the underlying structure and spatial scales of the nuclei of our target galaxies, we have produced simple models of their light distributions by fitting the data with combinations of point sources plus extended components with simple analytical forms. Our approach is motivated by the need to extract information from objects which are only slightly resolved and to contend with the undersampling of the PC2. Both of these effects limit the ability to uniquely reconstruct brightness distributions of small angular size sources, such as nuclei (e.g., Wildey 1992; Bouyoucef {\it et al.}\ 1997). For non-active or weakly active nuclei, our models are motivated physically by the structure of the Milky Way nucleus (e.g., Mezger {\it et al.}\ 1996) as it would appear at larger distances. Our models are also appropriate for star cluster nuclei hosting AGNs (e.g., Norman \& Scoville 1988, Perry \& Williams 1993). Our modelling approach has the advantage of requiring a minimum number of free parameters, while at the same time yielding measures of the characteristic sizes of the nuclei and limits on their central luminosity densities. Because our approach does not rely on deconvolution methods, we avoid the addition of noise and spurious structures at small radii that can be introduced by such techniques (e.g., Michard 1996). Deconvolution is especially problematic for undersampled PC2 data, where the solution is strongly dependent upon the pixel phase of the adopted PSF (see Hasan \& Burrows 1995). Moreover, if the structures of nuclei are discontinuous at small radii, and discreet point-like features are present near their centers, such information would not necessarily be recovered through deconvolution techniques (see Sams 1995). Finally, our modelling approach allows us to locate off-center or unsymmetrical features in our nuclei. The point source contribution for our fits was modelled using the Tiny Tim Version 4.1 software (Krist 1996) to produce a PSF for the PC2 for the appropriate filter, approximate $B-V$ color, and the spatial position on the PC2 CCD. The PSF was subsampled by a factor of 3 and then interpolated in order provide finer shifting and thus improve our ability to align the model PSF with the observations. In all cases jitter corrections were negligible. For the extended component of our fits, we considered several different analytical forms for the brightness distribution (see Table~3). Note that for $\gamma$=2.0, Model~(4) becomes the standard Hubble-Reynolds raw (Reynolds 1913; Hubble 1930). For $\gamma$=0.25, Model~(6) reduces to the de Vaucouleurs $R^{\frac{1}{4}}$ law (de Vaucouleurs 1948), and in the case of $\gamma$=2.0, Model~(8) is an isothermal sphere (also known as a ``modified Hubble'' or ``analytical King model''; see Binney \& Tremaine 1987). Model~(7) (King 1962; 1966) also reduces to an isothermal sphere in the limit $c\equiv log(r_{t}/r_{c})\rightarrow\infty$. As we discuss below, more complex models with additional free parameters cannot be adequately constrained by the present data. For each of our target galaxies we attempted fits to the nuclei using models of three categories: (a) pure PSFs; (b) pure extended models of each of the types listed in Table~3; (c) combinations of a PSF and extended models. For cases (b) and (c), it was necessary to convolve the extended component model with our Tiny Tim model of the intrinsic PSF of the optics and to account for the effects of pixel scattering on its observed appearance. We accomplished this transformation by subsampling the generated extended model component, convolving it with the central 9$\times$9 pixels of the subsampled Tiny Tim PSF, rebinning to normal sampling, and finally, convolving the extended model with the pixel scattering kernel. \subsection{Fitting Procedure} Our fitting procedure began with inputting initial guesses for the various free parameters in the models. Any or all of the free parameters listed in Table~3 were allowed to vary during fitting. In addition, we could also allow the peak intensity of the PSF ($I_{PSF}$) and the fractional pixel shift of both the PSF center, $(x_{c},y_{c})_{PSF}$, and of the extended component center, $(x_{c},y_{c})_{ext}$, to vary. With our software, fitting was accomplished using a two-dimensional non-linear least squares technique based on the CURFIT program of Bevington (1969). This method employs the Marquardt (1963) algorithm, which combines a gradient search and an analytical solution derived from linearizing the fitting function. Individual pixels were weighted using Poisson statistics. In all cases, a $\chi^{2}$ goodness-of-fit criterion was used to evaluate the quality of the models. For each fit, we show a 2-D representation of goodness-of-fit as: Merit = (Data $-$ Fit)$^{2}~\times$ Weight. These are a useful tool for comparing the goodness of fit between different nuclei, since due to normalization uncertainties, the absolute values of the derived $\chi^{2}$ measures are difficult to compare directly. Outputs from our fits included values for the free parameters for each model from Table~3, as well as for $I_{PSF}$, $(x_{c},y_{c})_{ext}$, $(x_{c},y_{c})_{PSF}$, the background level, and the integrated fluxes of the PSF and the extended component contributions to the final fit. To estimate the sensitivity of our fits to effects of time variability in the PSF and camera focus, to mismatches between the true PSF and our model PSF, and to the relative strengths of the underlying PSF and extended components, we generated test images composed of an {\it observed} $F814W$ PSF with different focus values added to a model extended component. For the extended model we adopted a generalized Hubble-Reynolds law [Model~(4)] with $r_{0}'$=1.0 or 1.5 pixels and $\gamma$=2.0 or 2.5. We tested focus offsets of up to 5$\mu m$ (i.e., 1/20$\lambda$), which is the maximum amount of defocus we expect to see in real images (Biretta {\it et al.}\ 1996). We then attempted to fit these test images with our fitting software using a nominal Tiny Tim PSF in perfect focus, plus a generalized Hubble-Reynolds extended component. We repeated this procedure on test images where the relative strength of the PSF component $I_{PSF}$ to the extended component $I_{0}$ ranged from 10 to 0.35. Regardless of focus, we found the fits output by our modelling software to underestimate the value of $I_{PSF}$ by $\sim$3-5\%. $I_{0}$ was accurately determined when $I_{PSF}$/$I_{0}$=0.35 but was overestimated by up to 33\% in the case of $I_{PSF}$/$I_{0}$=10. Errors on $\gamma$ range from 0-10\%, depending upon both the ratio $I_{PSF}$/$I_{0}$ and the true $\gamma$ of the test image. Finally, for the in-focus test images, $r_{0}'$ was systematically underestimated by 2-6\%, while in the de-focused test images, $r_{0}'$ was overestimated by 2-26\%, depending on $I_{PSF}$/$I_{0}$ and the true underlying $r_{0}'$ of the test image. In the discussions which follow, we use the above results as partial guidelines for estimating the uncertainty of the fit parameters derived from our modelling technique. As a second test of the accuracy of the PSF models produced from Tiny Tim, we attempted to use appropriately scaled versions of the Tiny Tim models to subtract field stars from our observed image frames. We found the Tiny Tim models to accurately subtract the stars to within $\sim$2$\sigma$. These residuals are consistent with those expected from the effects of Poisson noise and the large-angle scattering in the WFPC2 camera that is not modelled by Tiny Tim. We adopt the Tiny Tim model PSFs for our nuclear fits rather than using an observed PSF from each frame, since the observed point sources on our frames are of rather low signal-to-noise. In addition, because the PC2 PSF is position-dependent, the adopted PSF must be interpolated to match the location of the observed nucleus. This can be done with much greater accurately with a subsampled Tiny Tim PSF than with an observed PSF. From our tests we deduced the optimal extraction box size for fitting to be 31$\times$31 pixels. This size included the bulk of each nucleus' signal, while excluding most of the light from the large-angle scattering halo. Because M33 is brighter and better-resolved (and hence less sensitive to uncertainty in the PSF halo) we extended its box size to 61$\times$61 pixels. Modest changes in box size caused maximum changes of a few percent in our derived fit parameters. \subsubsection{Pure PSF Fits} To assess how well we had resolved our nuclei, we attempted fits to the nuclei of all program images using a pure point source (i.e., a Tiny Tim PSF). We find that {\it none} of the nuclei of our target galaxies can be adequately fit by a pure PSF. Fig.~2 illustrates this for the most distant nucleus in our sample, ESO~359-029. Clearly a PSF properly scaled to fit the profile wings grossly overestimates the central intensity. The mismatch is even more extreme in the other 3 cases. This implies that {\it we have partially resolved the nuclei of all four of our program galaxies}. For NGC~4395 and M33, the resolution is evident even from visual inspection of the images. \subsubsection{Pure Extended Component Fits} As a second step, we attempted to fit all four of our program nuclei using simple analytical brightness distribution models of the forms given in Table~3. We found that several of these models [(1), (3), (4), (6), (7), (8), and (9)] could adequately fit the outer wings of the nuclear brightness profiles, but none was adequate to fit the data over the full range of $r$ for any of the nuclei. In all instances the models severely underestimated the flux in the central regions of the nuclei. Fig.~3 shows this is the case even for the best resolved nucleus in our sample, M33. \subsubsection{Fits Including Extended Components Plus a PSF} To better assess the nature of the light distributions of our nuclei over the full range of $r$, our final series of modelling attempts consisted of fitting with combinations of a point source and an extended component from Table~3. Our goal was not only to try to reproduce the brightness profiles of each individual nucleus, but to see if we could find a single, unified characterization of all of the nuclei that would allow meaningful comparisons between them. For this reason, we began with our best-resolved case, M33. We first fit the M33 nucleus in the $F814W$ frame using combinations of a PSF and several of the ``classic'' brightness distributions given in Table~3. The exponential disk+PSF model fit the wings of the profile adequately, but overestimated the peak central brightness. The Vaucouleurs $R^{\frac{1}{4}}$ law+PSF model was a poor fit for almost the full range of $r$, and significantly overestimated the peak brightness level. The Hubble-Reynolds law+PSF combination roughly fit the shape of the profile for most values of $r$, but underestimated the peak central intensity. By far the best fit came from the King law+PSF fit, which provided an excellent model of the brightness profile for all values of $r$ and correctly reproduced the central intensity (Fig.~4). For this model we found the best fit occurred as $c\rightarrow\infty$---i.e., the best-fit King model tended toward the case of a simple, isothermal sphere [Model (9)]. Using more generalized fitting formulae with additional free parameters [e.g., Models (4), (6), and (8)] only marginally improved our fits. One can trade off between $\gamma$ and the characteristic radii of the models to produce a family of fits, none of them unique (see also KM). In general, the exponent $\gamma$ becomes larger as the characteristic radius ($r_{0}'$, $r_{e}'$, or $r_{c}'$) increases. Thus while they do reproduce the data adequately, these sorts of generalized fits have no obvious physical meaning, and do not permit useful comparisons between the different nuclei. Nonetheless, one important result did emerge: for the modified King model we found $\gamma\rightarrow 2.0$---i.e. this more general form again tended toward the case of a isothermal sphere. Together these results suggest that an isothermal sphere is an excellent approximation to the outer brightness distribution of the M33 nucleus, while at {\it HST}\ resolutions, the center of the nucleus is indistinguishable from a point source. Because our goal is to fit the observations with a useful and reproducible model rather than exploring all possible model fits (see also KM), we adopt for its simplicity and minimal number of free parameters, the isothermal sphere+PSF (hereafter IS+P) model as an analysis tool for all of our program nuclei. The emergence of the IS+P model for our program nuclei suggests that {\it to the resolution limit of our images, the central stellar density continues to increase in all four of our program nuclei.} Although we cannot conclusively rule out other classes of models from our present data (cf. L98), the IS+P model provides a useful and physically motivated characterization of our program nuclei. Below we present a more detailed analysis of the WFPC2 images of each of our nuclei, including results derived from IS+P model fitting. For each case, we interpret our findings by incorporating results from the present work as well as prior results from the literature. \section{Properties of the Target Nuclei} \subsection{The M33 Nucleus} Because of the proximity of M33, its very small bulge, and its moderate inclination, its nucleus can be effectively studied from the ground. Spectra by van den Bergh (1976), Gallagher {\it et al.}\ (1982), and spectral synthesis models of O'Connell (1983) and Schmidt {\it et al.}\ (1990) show that a range of stellar ages, including stars younger than about 1~Gyr, exist in the nucleus. Massey {\it et al.}\ (1996) deduced from FUV$-$NUV colors that the young population has a color consistent with a small group of \mbox{He\,{\sc i}}\ emission stars like that seen at the Galactic Center. High resolution ground-based imaging and spectroscopy by KM indicated the nucleus is very compact with a core radius $r_{c}<$\as{0}{10} ($<$0.4 pc), has a stellar velocity dispersion of $\sigma_{\star}$=21$\pm$3~\mbox{km s$^{-1}$}, $\frac{{\cal M}}{L_{V}}$$\sim$0.4, and a maximum central black hole mass of 5$\times$10$^{4}$\mbox{${\cal M}_\odot$}. Despite the lack of a very massive central black hole in the M33 nucleus, there are some indications of mild activity in the form of [\mbox{N\,{\sc ii}}] optical emission lines (Rubin \& Ford 1986), possible short- and long-term optical variability (Lyutyi \& Sharov 1990) and a moderately luminous ($\sim$10$^{39}$~ergs~s$^{-1}$), variable hard X-ray source (Schulman \& Bregman 1995; Dubus {\it et al.}\ 1997). Observations of the center of M33 with the WF/PC-1 Planetary Camera on {\it HST}\ were presented by Mighell \& Rich (1995). They measured a $V-I$ color-magnitude diagram for surrounding field stars which shows a broad red giant branch tip composed of stars with ages $\geq$1.7~Gyr, younger red supergiants, and a main sequence including stars with ages of $\sim$0.1~Gyr. We emphasize that M33 cannot be assumed to be simply a massive version of an ordinary globular cluster because its characteristics and environment are much more complex. The WFPC2 images of the M33 nucleus that we present here were also recently analyzed by L98. These authors found the M33 nucleus to be centrally peaked, which they interpret as a continuous increase in mean stellar density towards a central `cusp'. They also noted that the nucleus becomes somewhat bluer in color at small radii. L98 interpret this combination of properties as possibly being due to the presence of binary star merger products in the post-core collapse nuclear star cluster, as previously suggested by KM. L98 use a deconvolution analysis to derive a fit to the nuclear light profile of M33. Their fit is a steep power-law profile for \as{0}{05}$<r<$\as{0}{2} with a somewhat shallower central cusp or core with $r_{c}\approx$\as{0}{034}. They place a further limit on the maximum black hole mass of $M_{BH}<$2$\times10^{4}$\mbox{${\cal M}_\odot$}, and derive a central luminosity density of $\sim$5$\times$10$^{6}$\mbox{${\cal L}_\odot$}~pc$^{-3}$. A close inspection of the WFPC2 images of the M33 nucleus reveals several additional interesting features. In agreement with L98, we find the nucleus is clearly elongated, especially in the $F555W$ band, confirming the earlier suggestion of KM. We locate the major axis at a position angle (PA) of 18$^{\circ}$, in excellent agreement with the value of PA=17$^{\circ}$ determined by L98. At a PA of roughly -20$^{\circ}$ we see evidence of a radial, jet-like feature in both the $F555W$ and the $F814W$ frames (Fig.~5a \& b). This feature is roughly \as{0}{5} long, and contains two bright knots. In the present data, the structure of these two knots are both consistent with point sources to within errors, but their location, relative orientation, and colors are intriguing. The centroids of both of these bright spots lie along a radial line pointing directly toward the center pixel of the nucleus. The knots have absolute magnitudes (corrected for Galactic extinction) of $M_{V}$=$-$3.99 (top) and $M_{V}$=$-$5.21 (bottom), respectively, and there are no other discrete sources of similar luminosity in the outer parts of the nucleus or its surroundings. If these are single stars, they are rather luminous, and it is difficult to explain their highly disparate colors. The upper source is very red [($V-I$)$_{o}$=1.92] while the lower source is very blue [($V-I$)$_{o}$=$-$0.22] (see Fig.~6, discussed below). Even if the sources we see are luminous stars, their alignment with the nucleus center raises the possibility that they may be associated with a jet or outflow of some sort. Further investigation of these features is clearly desirable. If the M33 nucleus does harbor a miniature jet, this would provide evidence that M33 is indeed a low-level AGN. We attempted fits to the WFPC2 images of the M33 nucleus using our modelling software, as described in Sect.~3.2.3. Our IS+P model reproduces the light profile of the M33 nucleus quite well, aside from small systematic errors due to the slight ellipticity of the nucleus. The mismatch between our circularly symmetric models and the slightly elliptical nucleus of M33 results in a symmetric residual pattern upon subtraction of the model (Fig.~4 \& 7). Our derived core radius for the IS component of the M33 nucleus from the $F814W$ image is $r_{c}$=\as{0}{11}, which is in good agreement with the measurements of KM and Mighell \& Rich (1995), both of whom found $r_{c}\approx$\as{0}{1}. As in the present work, KM also included a central point source to reproduce the compact central source or ``cusp'' in the brightness profile. Our IS model value for the core radius is an {\it upper limit} to the physical size of any core in the M33 nucleus. Our value is larger than the slope break radius derived by L98. These authors interpret the unresolved center of the nucleus as a density cusp, whose properties were measured via deconvolution. Our core radius is derived from an isothermal sphere fit to only the {\it resolved} portion of the underlying nuclear star cluster. Our resulting IS+P model magnitudes agree well with our aperture photometry (see Sect.~4.7). For example, the observed $F555W$ magnitude for a 25 pixel circular aperture is only 0.01 magnitude fainter than that derived from our model. We can derive a limit to the central luminosity density by assuming the flux in the model point source in the unresolved center of the nucleus emanates from a region whose size is less than one PC2 pixel. At the distance of M33, this corresponds to a volume with radius 0.092~pc, and the minimum $V$-band luminosity density for the M33 nucleus of 3.6$\times$10$^{7}$~\mbox{${\cal L}_\odot$}~pc$^{-3}$. The L98 model yields a lower central density for the M33 nucleus since it assumes a continuous model for the nuclear radial brightness profile. The color map of the M33 nucleus (Fig.~6) suggests this nucleus has a blue core, as previously inferred by KM, Mighell \& Rich (1995), and L98. This result must be interpreted with caution due to the mismatch between the {\it HST} PSFs at different wavelengths, but nonetheless, the spatial extent of the blue region on our color map ($\sim$3 pixels) is consistent with the slightly bluer nuclear core measured by L98 in their deconvolved image. From aperture photometry (see below), we measure a $V-I$=0.81 within a 3-pixel radius aperture, also consistent with Fig.~21 of L98. If real, this central concentration of blue light could represent a small cluster of young stars (e.g., \mbox{He\,{\sc i}}\ stars, as suggested by Massey {\it et al.}\ 1996), a miniature AGN, or a single blue supergiant. If such discreet sources are present, then the fit of smooth power law to the radial intensity distribution becomes uncertain, as discussed by Sams (1995). Similar behavior is also seen in the Milky Way's nucleus, but in this case it is due to a central concentration of luminous young stars (Krabbe {\it et al.}\ 1995, Mezger {\it et al.}\ 1996). The central young cluster in the Milky Way nucleus would have a core radius of about \as{0}{03} at the distance of M33, and therefore would appear as a blue point source superposed on an older, redder underlying cluster, similar to what we see in M33. We also note that some blue light is present in the outskirts of the M33 nucleus; consistent with the color profile shown in L98, we see a faint, moderately blue ring surrounding the periphery of the entire nucleus (see Fig.~6). It is clear that even low-mass nuclei such as that in M33 are more complicated than simple scaled-up versions of ordinary star clusters, and the radial intensity profile, especially in the central regions, may reflect more than the stellar density profile. Therefore, caution must be exercised in assessing the dynamical state of compact nuclei from brightness profiles alone (see KM, L98). Because M33 is sufficiently resolved, we can also explore its brightness distribution using isophotal fits. We determined the run of surface brightness with radius for the $F555W$ 10-$s$ exposure using the ELLIPSE task in IRAF. To match the slight elongation of the nucleus, we used a fixed ellipticity of $\epsilon$=0.15 at a position angle PA=18$^{\circ}$. The background was measured in the regions surrounding the nucleus, and a constant background was subtracted. The resulting observed radial brightness profile in the $F814W$-band is shown in Fig.~8. For comparison, we also fitted the M33 nuclear brightness profile using circular isophotes. On an azimuthally-averaged brightness profile, the results appear virtually indistinguishable from the elliptical fits, indicating the circular symmetry in the IS+P models should not be a major source of uncertainty. \subsection{The M33-at-a-Distance Nucleus} Since our other target nuclei are more distant than M33, our WFPC2 observations in these cases suffer more severely from limited angular resolution. However, it is possible to make some comparative measures of the size and radial run of intensity in the outer regions of the other nuclei. This information provides a basis for testing the hypothesis that other star cluster nuclei in small spirals are structurally similar to the nucleus of M33. It is useful, before undertaking this analysis, to first explore how the observed properties of the M33 nucleus would change if it were moved outside of the Local Group. As a test of the effects of distance on our model fits, we block averaged the $F814W$ image of the M33 nucleus by 5$\times$5 pixels to simulate its appearance at a distance of about 4~Mpc (hereafter ``M33-at-a-Distance''). The WFPC2 $F814W$ image was selected for this experiment because it is less affected by recent star formation or possible effects of weak nuclear activity, and so provides the best measure of the intrinsic properties of the underlying nuclear star cluster. The main source of uncertainty in this experiment comes from the fact that in a real galaxy, an unresolved source would have its flux distributed over the same number of pixels regardless of its distance; however in our test image, the central point source was binned 5$\times$5. Our results for the IS+P fits to the M33-at-a-Distance nucleus are presented in Table~4 and Fig.~9. These fits demonstrate a few basic points about the impact of resolution on the fitting process. The fraction of the total integrated model flux contained in the PSF increases from 3.4\% in M33 to 20\% in M33-at-a-Distance. The residuals from the IS+P fit also become much smaller, demonstrating our lessened ability to resolve underlying structure of the nucleus, while at the same time the derived core radius of the extended model component yields a value $\sim$1.3 times larger (in parsecs) than for the M33 nucleus at its proper distance. Thus we cannot measure the form of the central part of distant nuclei without knowing their radial intensity profile in advance. These effects are to be expected. With increasing distance, more of the nucleus' light falls within an unresolved point source, and fewer resolution elements lie across the resolved component. As a result, the extended brightness component of the nucleus must be separated from the wings of a strong PSF and cannot be constrained as accurately. Since the PSF also effectively removes additional light from the extended component near the center of the image, the best-fitting extended component will necessarily have a shallower brightness gradient at small $r$. Thus any derived core radius for the IS component of the intensity model will only be an upper limit to the intrinsic $r_{c}$. \subsection{The ESO~359-029 Nucleus} Sandage \& Fomalont (1993) first reported the presence of a point-like nucleus in ESO~359-029, but classified the host as an Im/dE,N ``mixed morphology'' galaxy, and argued that it is gravitationally bound to the Sbc spiral NGC~1532. However, CCD imaging by Matthews \& Gallagher (1997) emphasized the disky appearance of this faint galaxy. They classified ESO~359-029 as an Sd, and noted that its disk was small and diffuse, but very symmetric, with a semi-stellar nucleus centered on an ``island'' of slightly higher surface brightness than the surrounding disk. In addition, a recent high-resolution \mbox{H\,{\sc i}}\ spectrum by Matthews {\it et al.}\ (1998) shows that ESO~359-029 is clearly rotationally dominated, that the velocity width reported by Sandage \& Fomalont (1993) was underestimated, and that the \mbox{${\cal M}_{{\rm H\hspace{0.04 em}\scriptscriptstyle I}}$}/$L_{V}$ ratio of this galaxy (0.41 in solar units) is entirely normal for extreme late-type spirals, while being a factor of 10 higher than typical dE or dwarf spheroidal systems (e.g., Oosterloo {\it et al.}\ 1996 and references therein). For these reasons we believe ESO~359-029 is best classified as an extreme late-type spiral of type Sd or Sdm. We note that ESO~359-029 is physically smaller than the other 3 galaxies in the present sample, and shows almost no hint of spiral structure. Like many extreme late-type spirals, its global properties are similar to those of an irregular galaxy (Matthews \& Gallagher 1997). This raises the interesting possibility that at least some nucleated dE galaxies may form when gas is stripped from an extreme late-type spiral during a close encounter with another galaxy (cf. Sandage \& Fomalont 1993). If extreme late-type spirals were the precursors of dEs, this would eliminate the difficulty of explaining their origins from irregulars, which are generally not nucleated (cf. Binggeli 1994). One way of testing such a scenario is to compare the nuclear properties of galaxies such as ESO~359-029 with those of dE nuclei observed at similar spatial resolutions. One candidate for such a comparison is the Local Group dE NGC~205. Its nucleus resembles those studied here in terms of its luminosity and size (Jones {\it et al.}\ 1996). A longslit spectrum taken by Matthews (1998) revealed that ESO~359-029 has detectable H$\alpha$ and [\mbox{N\,{\sc ii}}] emission over most of the optical extent of its disk, but no [\mbox{S\,{\sc ii}}] was detected. The rotation curve of this galaxy is shallow, slowly rising, and fairly linear, but shows an abrupt reversal near the nucleus, with a semi-amplitude of +25~\mbox{km s$^{-1}$}\ and -13~\mbox{km s$^{-1}$}\ on the approaching and receding sides of the nucleus, respectively. Thus there appears to be a kinematic signature of a compact, massive concentration at the center of this tiny spiral. The results from IS+P model fits to the ESO~359-029 nucleus are presented in Table~4 and Fig.~10 \& 11. ESO~359-029 is satisfactorily fit in both the $F450W$ and the $F814W$ bands by the IS+P model. We see only a very weak, slightly elongated residual in the $F814W$ frame and essentially no discernible residuals in the $F450W$ fit. The core radii we derive for the resolved component of the ESO~359-029 nucleus from our IS+P model fits are the largest of the four galaxies in our sample, although ESO~359-029 is also the most distant of the four and we emphasize our $r_{c}$ values are only upper limits (see Sect.~4.2). In the blue band 38\% of the model flux lies in the point source component of the fit, suggesting we have only marginally resolved this nucleus. In a manner similar to the case of M33, we derive a minimum central $B$-band luminosity density for ESO~359-029 from our IS+P model of 2.9$\times$10$^{4}$\mbox{${\cal L}_\odot$}~pc$^{-3}$ within a volume of radius 1.10~pc. Finally, we note that in spite of its faintness (and hence seemingly comparatively low mass), the nucleus of ESO~359-029 is clearly very compact. The nucleus of ESO~359-029 appears to share basic similarities in its properties with the other compact star cluster nuclei in our sample, hinting that the properties of compact star cluster nuclei are at least to some degree independent of the size and luminosity of their host galaxies. \subsection{The NGC~4242 Nucleus} In the atlas of Sandage \& Bedke (1994) the morphological appearance of NGC~4242 is very similar to that of NGC~4395. Van den Bergh (1995) drew attention of the prominent semi-stellar nucleus visible in that image. However, unlike the case of NGC~4395, there are no spectral signatures of activity in this nucleus. Ho {\it et al.}\ (1995a) published an optical spectrum of the NGC~4242 nucleus showing weak emission lines, and Heckman (1980) failed to detect a compact nuclear radio source down to a limit log$(L_{6 cm})<18.91$~W~Hz$^{-1}$. The lack of discernible FIR emission from this nearby spiral by {\it IRAS} and it relatively weak global radio continuum flux (Gioia \& Fabbiano 1987) both suggest a low rate of global star formation throughout the disk. Visual inspection of our images of the nucleus of NGC~4242 (e.g., Fig.~1c) reveals very faint, extended ``fuzz'' around the main bright nucleus. This is most evident in the $F814W$ image. Results from IS+P model fitting are presented in Table~4 and Fig.~12 \& 13. From this model we derive a minimum $B$-band central luminosity density for this nucleus of 1.0$\times$10$^{5}$~\mbox{${\cal L}_\odot$}~pc$^{-3}$ within a volume of radius 0.815~pc. We find the NGC~4242 nucleus is unlike the other nuclei in our sample in two important ways. First, this is the only one of the four nuclei that shows significantly more structure in the $F814W$ images than in the $F450W$ image. Our IS+P model provides a good fit to the $F450W$ brightness distribution, but in the $F814W$ band, this model cannot properly reproduce the flux distribution in the central few pixels, and leaves an oval-shaped residual pattern roughly \as{0}{5}$\times$\as{0}{3} across (see Fig.~13). This reveals that this nucleus in not circularly symmetric, and that it appears to contain a miniature bar-like feature near its center. We see marginal evidence for a similar feature in the $F814W$ image of ESO~359-029, but it is significantly less pronounced. This may be due to an intrinsic difference, or to the poorer spatial resolution in the ESO~359-029 images. The second difference between NGC~4242 and the other nuclei is that NGC~4242 is the only one of the three non-active nuclei where the point source contribution to our model fits is larger in $F814W$ than in the bluer image. \subsection{The NGC~4395 Nucleus} Filippenko \& Sargent (1989) were the first to present evidence that the compact nucleus of NGC~4395 is the lowest luminosity Seyfert~1 known. These authors discovered that the H$\alpha$ emission line in the NGC~4395 nucleus has a broad component, and that the [\mbox{O\,{\sc iii}}] emission is much stronger than that of H$\beta$. Further support for the presence of an AGN came from a UV spectrum taken with the Faint Object Spectrograph on the {\it HST} by Filippenko {\it et al.}\ (1993) which showed a flat far-UV continuum with strong, high ionization emission lines. No P-Cygni profiles were found in the UV, such as would normally be present from winds in massive OB stars. In addition, Filippenko {\it et al.}\ found the nucleus to be spatially extended in observations made with the original Planetary Camera through the $F502N$ filter (a narrow-band \mbox{O\,{\sc iii}}\ filter), while the continuum $F547M$ filter showed only a point source. Sramek (1992) measured a non-thermal radio source with about the luminosity of the Cas~A Galactic supernova remnant at the position of the NGC~4395 nucleus. Finally, Lira \& Lawrence (1998) have recently reported the nucleus is a variable X-ray source. Taken together, the spectral characteristics of the NGC~4395 nucleus appear to be more consistent with a standard accretion-powered AGN rather than a compact starburst. With our new WFPC2 data, we have resolved the nucleus of NGC~4395. This nucleus has complex internal structure. In both the $F450W$ and $F814W$ bands, some degree of elongation is visible, and a faint halo of irregular ``fuzz'' can be seen surrounding the brighter core of the nucleus. Because the NGC~4395 nucleus is less structurally complex in the $F814W$ frame, we modelled that image first. The resolution of the nucleus in $F814W$ is illustrated in Fig.~14, where we have subtracted a scaled PSF model to approximately fit the the central intensity. This leaves behind an extended, slightly elliptical residual pattern which contains about half of the light. A complete fit to the nucleus can be made with good accuracy using our IS+P model; this produces a point source and extended object centered at the same position to within better than 0.1 of a PC2 pixel ($\approx$5~milliarcsec). As with our other sample nuclei, the outer regions of the nucleus of NGC~4395 are reasonably fit by an IS+P model (Fig.~16). However, for NGC~4395 an asymmetric, bipolar-like pattern emerges in the residuals. The luminosity of the IS component of NGC~4395's nucleus in the $I$-band is $M_{I}\approx -$10.1 (see Sect.~4.7), similar to the resolved nuclear cluster components in the other galaxies in our sample. The luminosity and structural parameters of the NGC~4395 nucleus that we measure from the present data are consistent with NGC~4395 containing a normal star cluster nucleus which is currently hosting nuclear activity. However, final confirmation of this picture will require high angular resolution spectra. The NGC~4395 nucleus looks considerably different in the $F450W$ filter. Two of the central pixels of the $F450W$ image were saturated in the long exposure, hence these were replaced by values derived by scaling from our short exposure (see Sect.~2). Visually inspecting the $F450W$ image, we see that the nucleus of NGC~4395 is asymmetric and elongated; it can be described as a point source superimposed on an elongated, somewhat irregular structure. We have selected the contour levels in Fig.~16 to attempt to illustrate this point; note the displacement of the point source from the outermost isophote. Furthermore, unlike the other nuclei, where the centers of the PSF and IS components to the fits coincide to within a fraction of a pixel (see Table~4), the centroids of the PSF and IS components for the $F450W$ image of NGC~4395 are offset by 1.2 pixels, and the centroid of the IS component differs from that found in the $F814W$ image by nearly 2 pixels. The luminosity in the point-source component is considerably higher than would be expected from the effects of distance alone (based on our M33-at-a-Distance experiment), arguing that this nucleus truly contains a very compact source near its center, consistent with other AGNs. A fit to the light distribution in the $F450W$ frame using our IS+P model leaves a bipolar residual pattern similar to that in the $F814W$ frame, but much more pronounced, and containing an even larger fraction of the total nuclear light (Fig.~17). From our IS+P model we place a limit on the minimum $B$-band luminosity density in the NGC~4395 nucleus of 9.6$\times$10$^{6}$~\mbox{${\cal L}_\odot$}~pc$^{-3}$ within a volume of radius 0.283~pc. Further hints of the complex structure of the NGC~4395 nucleus can be gleaned from an examination of a $F450W-F814W$ color map (Fig.~18). We see overall the nucleus is fairly blue, with a compact central source that appears to be somewhat bluer than the surrounding regions. There is also a short, bright blue arc ($\sim$\as{0}{4} long) offset just a few pixels from the nucleus center. This and the prominent extended blue plume visible to the right of the nucleus are discussed in detail below. As with the $F814W$ frame, we examine the residuals left after subtracting only the point source component of our model (Fig.~19). From this we see evidence of a slight elongation at a PA=255$^{\circ}$ (at ``2 o'clock'') in the $F450W$ image. No analogous feature was seen in the $F814W$ image (Fig.~14). The elongated structure is only weakly visible, but if the feature were simply an artifact of an error in the PSF, we would expect to see a reflected symmetry in this structure, and we do not. The stronger evidence of the reality of these features is that it appears to correspond to a feature in our contour plot (Fig.~16) and to a distinct blue arc in the color map shown in Fig.~18. In the color map, the arc-like feature is located about \as{0}{25} ($\sim$3~pc) from the center of the nucleus. The structure of this arc and its extremely blue color are consistent with it being produced by emission lines from ionized gas (see below). These asymmetric blue structures we see in the NGC~4395 nucleus are most readily understood if the light is due to gas rather than stellar emission; dynamical time scales within nuclei are extremely short and thus any azimuthal irregularities in the distributions of stars will rapidly disappear in much less than 1~Myr. A stronger confirmation of this picture comes from the consistency of our observed asymmetry in the NGC~4395 nucleus with that found by Filippenko {\it et al.}\ (1993) in their $F502N$-band observations. This agreement suggests that the inner and outer filaments are due to regions where the H$\beta$ and [\mbox{O\,{\sc iii}}] emission lines are especially intense. Using the WFPC2 efficiencies tabulated by Biretta {\it et al.}\ (1996), we find that the blue arc-like feature produces a flux in [\mbox{O\,{\sc iii}}] of $\sim4.5\times10^{-14}$~ergs~s$^{-1}$~cm$^{-2}$, which is about 1/4 to 1/3 of the [\mbox{O\,{\sc iii}}] flux measured from ground-based spectra by Ho {\it et al.}\ (1997a), and corresponds to a luminosity of roughly 1$\times10^{4}$~\mbox{${\cal L}_\odot$}. The $F450W$ image of NGC~4395 contains a second, large emission plume that is clearly resolved from the main nucleus. It is approximately aligned with the inner blue arc discussed above. This plume also has the blue colors expected for an emission line source, and is only faintly visible in the $F814W$ image. The outer plume has a wispy, filamentary structure, and extends about 1$''$ ($\sim$20~pc) in projected distance from the nucleus, with its main structure centered near PA=265$^{\circ}$. Near PA=270$^{\circ}$, it makes a 90$^{\circ}$ turn and joins onto a fainter region at PA=280$^{\circ}$. We detected an integrated flux from the outer plume is 44 photons~s$^{-1}$, corresponding to corresponding to a flux in [\mbox{O\,{\sc iii}}] of 4$\times 10^{-14}$~erg~s$^{-1}$~cm$^{-2}$, or $L(\mbox{O\,{\sc iii}})\ge 1.5\times10^{38}$~erg~s$^{-1}\approx 1\times10^{4}$\mbox{${\cal L}_\odot$}. Thus this emission line plume, together with the arc discussed above, contain 1/2-2/3 of the [\mbox{O\,{\sc iii}}] flux found by Ho {\it et al.}\ (1997a) in their spectra taken with a 1$''\times$4$''$ entrance aperture. Since Ho {\it et al.}\ (1997a) quote an uncertainty of $\pm$20\% for their measurement, our values are in rough agreement. We therefore believe that these two features comprise the bulk of the narrow emission line region around the Seyfert nucleus of NGC~4395. Our observations of the structure and luminosity of the NGC~4395 nucleus and its surroundings are consistent with the behavior of more luminous AGNs. Normal Seyfert galaxies often show ``ionization cones'' along which highly ionized material is found (e.g., Pogge 1989a,b; Wilson {\it et al.}\ 1993; Boksenberg {\it et al.}\ 1995; Mulchaey {\it et al.}\ 1996; Schmitt \& Kinney 1996). This phenomenon is interpreted as blocking or collimation of radiation from the vicinity of an accreting black hole. The presence of two distinct filaments along approximately the same position angle from the nucleus gives strong evidence for the presence of an ionization cone in NGC~4395. This feature, in combination with the spectrum and brightness of the probable [\mbox{O\,{\sc iii}}] line emission, provide further support to the presence of an AGN within the central parsec of the NGC~4395 star cluster nucleus. These factors strengthen arguments against a stellar origin (e.g., the ``Warmer'' model of Terlevich \& Melnick 1985) for the activity in NGC~4395 (see also Filippenko {\it et al.}\ 1993). \subsection{Comparison of the Structural Properties of the Nuclear Star Clusters} The emergence of the isothermal sphere as a suitable approximate model for the outer brightness profiles of our nuclei suggests that the nuclei of all of our sample galaxies harbor true compact nuclear star clusters, not simply small, nuclear \mbox{H\,{\sc ii}}\ regions. In contrast to the isothermal sphere, the more general King law (King 1962,1966) includes a finite value for the cutoff or ``tidal'' radius $r_{t}$ (see Eq. 5, Table~3). What does it mean then that our present nuclei are better fit as $r_{t}\rightarrow\infty$ (or equivalently, $c\rightarrow\infty$) than with the more general King Law? One possibility is that we are unable to measure $r_{t}$ due to our limited field of view and confusion from the background light of the galaxy. However, in the center of a rigidly rotating galaxy, the tidal radius loses its usual physical interpretation, and the nucleus may simply blend into the disk, as discussed by KM. The result would then be a very slow drop-off in stellar density from a nuclear star cluster in galaxies with shallow central gravitational potentials; stars will only become unbound from the cluster at radii where the effects of galactic rotation begin to dominate. A comparison of the properties we derive for our sample nuclei from IS+P model fitting reveal both some interesting similarities and some important differences. The fit results yield two cases (M33 and ESO~359-029) where a larger percentage of the integrated model fluxes lie in the point source component at bluer wavelengths. Since the intrinsic resolution is slightly better at shorter (bluer) wavelengths, this may reflect a real, underlying property of these nuclei---i.e., that they contain a compact, blue central source (or possibly a blue ``cusp''). In comparing the core radii we derive for the IS component of each of our nuclei (Table~4), we see some expected effects of distance. As predicted from our test fits to M33-at-a-Distance, the fraction of the total flux in the central PSF increases from M33 to NGC~4242 at a distance of 7.5~Mpc and again for ESO~359-029 at 10.2~Mpc. If the overestimation of the core radius that we measured for M33-at-a-Distance can be assumed to scale roughly linearly with distance, we estimate that the star cluster components of all four of our sample nuclei have similar sizes. Despite these uncertainties, we note that the nuclei for which we have both $F450W$ and $F814W$ data, our derived core radii from the IS+P models are all $\sim$2 times larger in the $F814W$ images than in F450W. This result is independent of the differing fraction of the flux in the PSF component in the different galaxies. This provides a tantalizing hint that our inactive nuclei have underlying structural similarities---i.e. an outlying population of old or intermediate age red stars dominating the light at larger radii, and an increased blue luminosity contribution toward the center. This type of segregation is found in the Milky Way nuclear star cluster, although in the Milky Way nucleus the core radius of the older, red stars is $\sim$5-10 times that of the inner blue population (see Eckart {\it et al.}\ 1993, Rieke \& Rieke 1994, Krabbe {\it et al.}\ 1995). Alternatively, blue nuclear cores could result from dynamical processes, such as enhanced binary production leading to excess populations of blue straggler stars (KM; L98), or ultra-low luminosity AGNs. \subsection{Aperture Photometry of the Nuclei} We established in Sect.~4 that none of our program nuclei are fully resolved. Only the nucleus of M33, which can be traced over a diameter of about 7$''$ (28~pc), comfortably exceeds the size of our model of the WFPC2 PSF. However, in all cases the nuclei stand out from their surroundings on the WFPC2 images (Sect.~5), and thus aperture photometry is relatively straightforward. This is a major advantage of {\it HST} data, which typically offer at least ten times greater angular resolution than optical observations from the ground. Aperture photometry was carried out using standard software packages in IRAF. We selected aperture sizes to include approximately 90\% of the light from each nucleus. For the M33 nucleus we used radii of 50, 75, and 85 pixels (\as{4}{6}, \as{6}{9}, and \as{7}{8} diameters), 30 pixels radius (\as{2}{76} diameter) for NGC~4242 and NGC~4395, and 20 pixels radius (\as{1}{84} diameter) for ESO~359-029. Images were background-subtracted prior to photometering using background measures obtained from regions adjacent to the nucleus. Because our galaxies all lack bulges, the disk background light was comparatively weak relative to the nucleus. We derived magnitudes in the WFPC2 system as described in Holtzman {\it et al.}\ (1995b). These are given for each WFPC2 filter $i_{\lambda}$ by: \begin{equation} m(i_{\lambda}) = -2.5log(DN~s^{-1}) + 0.748 + ZP(m(i_{\lambda})) \end{equation} \noindent for observations made with GAIN=7. $DN$ is the data number, and the filter zero points, $ZP(m(i_{\lambda}))$, are given in Table~9 of Holtzman {\it et al.}\ (1995b). We adopted observed zero points for the $F555W$ and $F814W$ filters and the synthetic zero point for $F450W$.\footnote{Technically, magnitudes on the WFPC2 system are defined for a point source observed through a \as{0}{5} aperture; there is no standard convention for adapting this system to extended sources.} The resulting WFPC2 instrumental magnitudes for M33 were converted to $VI$ magnitudes following Holtzman {\it et al.}\ (1995b), while those of the other galaxies were transformed to the $BI$ system using new relationships derived by Holtzman (1998): $$I = m_{F814W} - 0.002(B-I) - 0.009(B-I)^{2}~~~~~(B-I)<1.3$$ $$I = m_{F814W} - 0.004(B-I) - 0.003(B-I)^{2}~~~~~1.3<(B-I)<3$$ $$B = m_{F450W} + 0.108(B-I) - 0.008(B-I)^{2}~~~~~(B-I)<1.3$$ $$B = m_{F450W} + 0.315(B-I) - 0.047(B-I)^{2}~~~~~1.3<(B-I)<3.0$$ \noindent Here $m_{F814W}$ and $m_{F450W}$ refer to the WFPC2 system magnitudes given by Eq.~1. Our $B$ and $I$ magnitudes are presented in Table~5. Past ground-based measurements can be used to check our results for the M33 nucleus. Photometry by Nieto \& Auri\`ere (1982) and by KM yielded $B=14.5\pm 0.1$ and $B=14.57\pm 0.07$, respectively. Combining our observed value of $V=13.80\pm0.05$ with ($B-V$)=0.68 measured by Walker (1964), we find excellent agreement between our new WFPC2 photometry and ground-based measurements, as well as with the WFPC2 measurements of L98. Our $V$-magnitude also agrees within errors with the value of $V$=13.82 measured by Sharov \& Lyutyi (1988). Absolute magnitudes in Table~5 are derived from the distances and extinction values given in Table~1. Our sample is not large enough to adequately explore possible correlations between host galaxy luminosity and the luminosity of the nuclei, although we do find the most luminous nucleus (that of M33) in the most luminous host galaxy and the least luminous nucleus (that of ESO~359-029) in the intrinsically faintest host galaxy. However, what is perhaps more intriguing is the {\it similarity} of the luminosities ($M_{I}\sim$-11$\pm1$) of all of the nuclei in spite of the range in luminosities of their host galaxies from $M_{V}=-15.1$ to $M_{V}=-18.2$. This demonstrates that nuclei of similar luminosities can be either active or non-active and raises the possibility that the nuclei share similar origins and long term evolutionary histories. In spite of the fairly small range in luminosities of our target nuclei, the spread in their colors is significant. The most luminous nucleus, that of M33. Interestingly, M33, while having the reddest nucleus in our sample, has the bluest nucleus in the Local Group (Sharov \& Lyutyi 1988). NGC~4242 and ESO~359-029 have intermediate colors, both somewhat bluer than typical globular clusters. Finally, the active nucleus in NGC~4395 stands out because of its extremely blue optical colors, making it nearly as luminous as the M33 nucleus in the blue. These blue colors argue against the suggestion of Carollo {\it et al.}\ (1997) that faint nuclear star clusters in quiescent disk galaxies are {\it old} stellar clusters, but rather hints that they could contain a mixture of stellar ages. Alternatively, binary mergers in old star cluster nuclei could yield a range in optical colors (see L98), as could the presence of low luminosity AGNs. Spectroscopy and spectral synthesis studies will be needed to better explore this issue. Such studies may also yield clues as to why a spread of nearly one magnitude in $B-I$ exists among the three non-active nuclei in our sample. One possible explanation for the observed color spread is that episodic star formation occurs in these types of nuclei (e.g., Firmani \& Tutukov 1994). All three galaxies have similar inclinations ($i\sim 50^{\circ}$), so reddening from extinction in the surrounding galactic disks is unlikely to contribute much to this spread. \section{Environments of the Nuclei} A striking characteristics of the nuclei in our sample is the visual appearance of each relative to its respective surrounding galactic disk. In all cases, the nuclei are ``naked''---i.e., they appear very distinct from their circumnuclear environment, and there is little evidence for significant quantities of dust or other material at the centers of these galaxies (see Fig.~1). While to some extent, the nuclei in the present sample were selected on this basis, we have confirmed this holds even at space-based resolutions. These nuclei thus represent density contrasts of roughly 1000:1 compared to their surroundings. This point becomes especially evident in the $F814W$ image of NGC~4395, where what appear to be two background galaxies can be seen through the disk, only tens of parsecs from the nucleus (Fig.~21; for contrast with luminous spirals, see the collection of {\it HST}\ nuclear region images of Carollo {\it et al.}\ 1997). Furthermore, for all of our sample galaxies, the continuum light from the galaxy disk is only weakly detected during the WFPC2 exposures, and we see no hints of spiral structure. While all the disks are actively forming stars as evidenced by the presence of H$\alpha$ emission, there is little central concentration of star formation in any of the galaxies (e.g., Wang {\it et al.}\ 1997; Gallagher \& Matthews unpublished). These properties of the nuclear regions of our sample galaxies set their nuclei apart from those in most other well-studied nearby spiral galaxies where the nuclei occur within an obvious bulge component, or the nuclei lie at the center of a spiral pattern (e.g., Pi\c{s}mi\c{s} 1987 and references therein; Carollo {\it et al.}\ 1997; Regan \& Mulchaey 1998) and where significant quantities of molecular gas and dust exist in the inner regions of the disk (cf., Carollo {\it et al.}\ 1997; Devereux {\it et al.}\ 1997). \section{Comments on Possible Formation and Evolutionary Scenarios for Compact Star Cluster Nuclei in Extreme Late-Type Spirals} A possible clue to the origin of star cluster nuclei in extreme late-type spirals comes from the observation of van den Bergh (1995) that nuclei are {\it not} seen in irregular galaxies, even when these galaxies are of similar luminosities to the objects studied here. Artyukh \& Ogannisyan (1991) also noted that compact nuclear radio sources (indicators of nuclear activity) are not generally found in irregular galaxies. This hints that somehow the presence of an organized disk is a requisite factor for the presence of a compact nucleus. The rotations curves of extreme late-type spirals tend to be almost linear to the last measured point in the optical galaxy and have very shallow velocity gradients (e.g., Goad \& Roberts 1981; Wevers 1984; Matthews 1998). This indicates that the centers of these galaxies are not particularly ``special'' places, and that they have relatively flat gravitational potential wells (see also Colin \& Athanassoula 1981). This point was also emphasized by KM and Filippenko \& Sargent (1989) for the cases of M33 and NGC~4395, respectively. By contrast, nucleated galaxies with luminous bulge components generally have steep density profiles with potentials pointing sharply towards their centers. There is some uncertainty whether the nuclei of our sample galaxies are even located precisely at the dynamical centers of these galaxies (see also Miller \& Smith 1992). Minniti {\it et al.}\ (1993) noted that the nucleus of M33 is displaced by $\sim20''$ from its small ``bulge'' component, and it is also displaced from the center of true neighboring disk isophotes (de Vaucouleurs \& Freeman 1970; Colin \& Athanassoula 1981). In addition, the velocity centroid of the nucleus of ESO~359-029 appears to be displaced by $\sim$7~\mbox{km s$^{-1}$}\ relative to the galaxy centroid (Matthews 1998). Taken together, the global and dynamical properties of nucleated extreme late-type spiral galaxies raise interesting new questions on the symbiosis between the nuclei and the parent galaxies and on the origin of these centralized mass concentrations in otherwise very diffuse disks Various models proposed for the formation of nuclei in early-type galaxies are problematic for the ``naked'' nuclei in extreme late-type spirals. Capuzzo-Dolcetta (1993) and Capuzzo-Dolcetta \& Vignola (1997) have discussed the possibility that the nuclei of some early-type galaxies may form from globular clusters sinking to the center of the galaxies via dynamical friction. However, the efficiency of dynamical friction is greatly reduced in galaxies with low central densities (e.g., Lin \& Tremaine 1983) making it doubtful that this scenario would be efficient in extreme late-type spirals. Finally, it appears that at least some extreme late-type spirals do not contain significant numbers of globular cluster systems (Matthews 1998). Another possibility is that the nuclei formed {\it in situ} from gas infalling onto their centers either due to starbursts (e.g., Firmani \& Tutukov 1994) or accretion of primordial gas clumps or small gas-rich galaxies (e.g., Loeb \& Rasio 1994). While it seems plausible that infalling gas may have provoked episodic new star formation and altered the stellar populations of an {\it existing} nuclear star cluster in our galaxies (e.g., Firmani \& Tutukov 1994; Tutukov \& Kr\"ugel 1995), it is still difficult to explain how gas infalling into shallow potentials like those in extreme late-type spirals could have formed the compact star clusters at their centers. Less efficient angular momentum transport and weaker gravitational potentials in extreme late-type spirals provides a natural explanation for why they appear not to have formed supermassive black holes at their centers, but it cannot account for the extraordinary compactness of the nuclear star clusters. This conundrum was pointed out by KM with respect to M33. Our new data reveal that M33 is not an anomaly and that whatever the conditions necessary for the formation of a compact nucleus in extreme late-type spirals, these conditions commonly occur in these galaxies. One possible means of delivering gas to galaxy centers to fuel activity or enhanced star formation is via a bar (e.g., Shlosman {\it et al.}\ 1990; Ho {\it et al.}\ 1995b; Ho 1996). Among our sample objects, NGC~4242 is generally classified as weakly barred, while none of the other galaxies, including M33 have obvious bars. Ho (1996) also reported that bars do not appear to enhance the probability of finding a compact nucleus in galaxies later than Sbc (see also MacKenty 1990; Ho {\it et al.}\ 1997b; Mulchaey \& Regan 1997). However, it has been pointed out that secondary nuclear bars $\sim$1~kpc in length may be needed to effectively transport material into the central parsec region of a galaxy (e.g., Shlosman {\it et al.}\ 1989; Friedli \& Martinet 1993). Such features have been detected in a few galaxies (Friedli {\it et al.}\ 1996), but we do not see any evidence of such bars in our program objects. We cannot however rule out that such structures may have existed in the past and aided the building of the observed nuclear star clusters. Sofue \& Habe (1992) suggested that late-type (Sc and Sd) galaxies may have been formed in low density environments and suffered at most weak tidal interactions during their evolution. This explains the lower typical masses of these systems and their lack of large bulges. In this model, early-type galaxies formed in denser environments and developed bulges via tidally-induced starburst events. If these interactions are also responsible for nucleus building and the fueling of AGNs (e.g., Loeb \& Rasio 1994), this also could explain the lack of evidence for supermassive black holes in extreme late-type spirals. The low-level activity in NGC~4395 may have been caused by a very weak interaction or the accretion of a small, gas-rich satellite that was not enough to cause a significant alteration in the appearance of the galaxy. In the rare cases, such as the galaxy 0351+026, where extreme late-type spirals are observed to harbor more powerful active nuclei (see Sect.~1.1), the tidal effects believed responsible for this activity (Bothun {\it et al.}\ 1982a,b) may irrevocably change the appearance of the host galaxy, thus explaining the rarity of such objects in the present epoch. An alternative picture is that if primordial massive black holes are common in galaxies (e.g., Silk \& Rees 1998), the presence of one in a galaxy may lead to ``puffing'' of the disk (e.g., Faber {\it et al.}\ 1997; Merritt 1998) and the formation of a bulge or spheroid; hence, by definition, none of these systems will be observed as extreme late-type spiral disks (cf. Matthews 1998). M33 is the only case where our present data can help to constrain a more detailed model of the dynamical evolution of the nucleus. Because of its small velocity dispersion, the predicted relaxation time of the M33 nuclear star cluster is only $t_{relax}\leq$2-3$\times10^{7}$~yr (assuming $r_{c}$=0.1~pc; Hernquist {\it et al.}\ 1991; KM). The colors of the M33 nucleus are also consistent with most of the stars being older than a relaxation time, so core collapse probably has occurred in M33. Our new upper bound to the the radius of any core in the nucleus of M33 strengthens this estimate of the relaxation time for simple models of core collapse (e.g., Spitzer \& Hart 1971), although the effects of stellar encounters, stellar evolution, and a range of initial stellar masses may still create order of magnitude uncertainties in this value (KM; L98). Our measured value of the approximate core radius for our IS model component of the nucleus of M33 (0.11~pc) also agrees with theoretical predictions of the core-collapse scenario (Hernquist {\it et al.}\ 1991). L98 emphasize that stellar collisions may have particularly important consequences in the M33 nucleus, which lacks a supermassive black hole to stabilize its structure. However, it is likely that the structure of the M33 nucleus is more complex than that of an ordinary star cluster, and the detailed stellar mass density profile in the center of the nucleus is therefore still uncertain. These issues can be most directly resolved by spectroscopic observations designed to measure any radial trends in stellar content, possible signatures of a miniature AGN, and the nuclear kinematics. \section{Discussion and Suggestions for Future Work} We have shown that the nuclei of our program galaxies exist in similar, low density galactic disk environments and have similar luminosities and spatial extents. All of these factors provide evidence (but not proof) that the four nuclei have shared common formation histories and are comprised of roughly similar underlying nuclear star clusters. We postulate that the color and morphological differences we observe among the various nuclei may be due to relatively short-lived evolutionary phases (e.g., episodes of enhanced star formation or nuclear activity) superposed on otherwise similar nuclear star clusters. M33's nucleus is the only one of our program objects for which the dynamical information exists, hence it is the only case for which a definite upper limit may be placed on the mass of a possible black hole at its center. We therefore know that any black hole in the nucleus of M33 must have a small mass ($M_{BH}<2\times10^{4}$\mbox{${\cal M}_\odot$}; KM; L98). Clearly similar measures for the other program nuclei are needed to further test the hypothesis that the other nuclei in our sample resemble the nucleus of M33. A velocity dispersion measure is of particular interest in the case of NGC~4395, since it can help to assess whether the activity in NGC~4395 is due to special properties, or whether it represents a transient phase that M33 may have experienced in its past. Filippenko (1992) has estimated that if the nucleus of NGC~4395 contains a black hole accreting at roughly the Eddington limit, the black hole mass would only be $\sim$100\mbox{${\cal M}_\odot$}. On the other hand, Filippenko points out that the formalism of Wandel \& Yahil (1985) applied to NGC~4395 yields an approximate black hole mass of 4$\times$10$^{4}$\mbox{${\cal M}_\odot$}. This estimate assumes that the observed cloud speeds in the NGC~4395 nucleus are produced by gravity. This black hole mass estimate is nearly identical to the upper limit derived for M33 by KM and L98, so it is consistent with the nuclei of these two galaxies harboring similar central massive compact objects. Further evidence for or against a ``unified'' picture for the nuclei of M33 and NGC~4395 could come from X-ray observations. Filippenko {\it et al.}\ (1993) quoted an unpublished X-ray luminosity of the NGC~4395 nucleus of 6.6$\times$10$^{37}$~erg~s$^{-1}$, which is consistent with the measurement obtained by Cui (1994). However, Lira \& Lawrence (1998) report a mean $L_{X}\sim9.7\times10^{37}$~erg~s$^{-1}$ (for D=2.8~Mpc) with a factor of 2 variability over a 15-day period. This is roughly a factor of 10 smaller than the X-ray luminosity of the M33 nucleus, but given the uncertainty in the distance to NGC~4395, and its slightly smaller total nuclear luminosity, it is conceivable that the X-ray luminosity of both the M33 and NGC~4395 result from similar mechanisms. However, the reported X-ray variability occurs on timescales significantly shorter than the 106-day X-ray variability reported for the M33 nucleus by Dubus {\it et al.}\ (1997). \section{Summary} We have analyzed WFPC2 Planetary Camera broad-band imaging observations of the compact star cluster nuclei of four extreme late-type spiral galaxies: M33, NGC~4395, NGC~4242, and ESO~359-029. All of these galaxies have diffuse, moderate-to-low surface brightness disks, low luminosities, little or no bulge component, and relatively weak central gravitational potentials. The nucleus of NGC~4395 is a previously known low-luminosity Seyfert~1; M33 has some signs of possible weak activity, while the other two nuclei are not known to be active. However, we have confirmed that all four nuclei appear to be true compact star cluster nuclei rather than simply small nuclear \mbox{H\,{\sc ii}}\ regions. All of the program nuclei are partially resolved with the Planetary Camera 2. The radial brightness profiles in all four nuclei cases is well fit by a combination of an isothermal sphere (IS) and a central point source (PSF) component, which we refer to as the ``IS+P'' model. Physically this model implies that the structure of the nuclei may consist of an underlying star cluster with an abrupt change in luminosity density near the center. This gradient may be due to a power law cusp in the stellar density profile, such as those associated with core collapse (e.g., L98), or could represent the presence of a point-like source such as an AGN, a compact group of young stars, or a single supergiant star. The existence of an unresolved center in all of our program objects indicates that the luminosity densities in the centers of the nuclei are increasing to the resolution limit of our data. In the case of M33, the IS+P model provides an equally good fit to the data as the continuous `nuker' power law model of L98. Our choice of the simple IS+P model is physically motivated by the presence of a compact central light concentration in the Milky Way's nuclear cluster (seen superposed on a smoother nuclear star cluster background), and the inferred presence of AGNs {\it within} compact nuclear star clusters (e.g., Norman \& Scoville 1988). A central subcluster of young stars like that in the Milky Way's nucleus would be at most marginally resolved by WFPC2 at the distance of M33, while an AGN would be unresolved (see Blandford 1990). The IS+P model therefore is on a firm physical basis for the NGC~4395 nucleus, where the point source component represents an extremely compact AGN. More complex models can provide equally good or slightly better fits to the observed nuclear brightness maps, but these have more free parameters and provide at most small improvements in the fit accuracy. Furthermore, the more complex models could not be uniquely constrained and hence do not supply physically insightful information. While the IS+P model is simplistic, it is also well-matched to the degree of resolution we can achieve for galaxy nuclei outside of the Local Group. We show that because of the effects of increasing distance, the core radii we derive for the underlying nuclear star clusters of our program galaxies are only upper limits. We estimate the severity of the effects of distance on our measurements by producing an artificial ``redshifted'' version of the M33 nucleus. From that experiment, we conclude that the sizes of the compact star cluster components of all four of our program objects are similar to within a factor of 4 and all have core radii of less than 1.8~pc in the $F814W$ band. The luminosities of all four nuclei are similar ($M_{I}\sim -11\pm$1) indicating that nuclei with similar luminosities can be either active or non-active. In spite of the modest range in optical luminosity, we see a spread of 1.81 magnitudes in the $B-I$ colors of our program nuclei, with the active nucleus of NGC~4395 being by far the bluest. This spread in color may be indicative of different evolutionary phases (e.g., level of nuclear activity or recent star formation) in otherwise structurally similar underlying nuclei. In spite of their compact sizes, the nuclei in our sample exhibit complex structure. M33 has slightly elliptical isophotes and shows the presence of a jet-like feature. This jet-like feature may be a signature of low-level activity in the M33 nucleus. NGC~4395 is even more structurally complex, especially in the $F450W$ frame, where it appears as a point source superimposed on elongated, somewhat irregularly-shaped isophotes, possibly resulting from ionized gas emission. In addition, there is a bright arc very near the nucleus ($r \sim$3~pc) which is likely due to [\mbox{O\,{\sc iii}}] emission, with $L$(\mbox{O\,{\sc iii}})$\approx1\times10^{4}$\mbox{${\cal L}_\odot$}. A larger filamentary structure is found along the same line of sight at $r \sim$20~pc, and is also likely to be a highly ionized emission region. This pair of structures seems to be analogous to the ``ionization cones'' seen in more powerful AGNs. These features, in combination with the luminous blue point source our model fits reveal, lend further support to a standard AGN model for the source of activity in the NGC~4395 nucleus (see also Filippenko {\it et al.}\ 1993). The similarity of the appearance of the NGC~4395 nucleus in the $F814W$ filter to the other nuclei in our sample suggests that it is a fairly normal star cluster nucleus which happens to be presently hosting activity. Stellar velocity dispersion measurements will help to confirm or refute this picture. NGC~4242 is the only one of the four program nuclei that appears more structurally complex in the $F814W$ frame than in the blue image. Our IS+P model fits uncover an elongated bar-shaped residual structure at the center of this nucleus. ESO~359-029 appears relatively symmetric in our data, but this is the most poorly resolved of our program nuclei. For the three nuclei for which we have $F450W$ data, we measure an IS model component core radii that are roughly twice as large in the $F814W$ band as in the $F450W$ band. The blue luminosity therefore appears to be centrally concentrated, a symptom that could result from a variety of processes. As in the Milky Way's central star cluster, the star clusters in our program objects appear to have outlying populations of old or intermediate-age red stars, and a larger mixture of stellar ages in their centers. The presence of miniature AGNs could also produce this trend, as would an excess population of `blue straggler' stars which could form via stellar mergers within dense nuclei (KM; L98). Various factors make the origin of compact star cluster nuclei in small spiral galaxies enigmatic. Our new observations confirm that even when viewed at very high angular resolutions, all four of our program galaxies appear to be ``naked''---that is they exist in extremely diffuse circumnuclear environments with a low-density surrounding stellar disk, no spiral arms or significant bulge, few regions of enhanced star formation, and little indication of dust. These environments are in stark contrast to those typical of brighter compact nuclei and more powerful AGNs. In addition, the central gravitational potentials of the host galaxies of our program nuclei appear to be very shallow. Thus it remains unclear how dense, compact collections of material formed at their centers. Furthermore, the morphologies of the host galaxies argue against their having been victims of major interactions (Matthews \& Gallagher 1997), and their low central densities mitigate against dynamical friction being efficient enough for a star cluster formed elsewhere to migrate to the galaxy center. \acknowledgements LDM has been funded by a graduate internship with the Wide Field and Planetary Camera 2 (WFPC2) Investigation Definition Team, which is supported at the Jet Propulsion Laboratory (JPL) via the National Aeronautics and Space Administration (NASA) under contract No. NAS7-1260. This work has been carried out as part of the WFPC2 Investigation Definition Team science programs. We thank D. M. Peterson for comments on an earlier version of this manuscript. \clearpage
\section{Introduction} At present all information on $\mathrm{CP}$ violation is consistent with the standard model (SM) of electroweak interactions and its Cabibbo-Kobayashi-Maskawa (CKM) mechanism of $\mathrm{CP}$ breaking \cite{ckm}. In this Letter, we address the question of whether or not current experimental data, in particular the new limit for the $B_s^0$ oscillation frequency, exclude the possibility of having a real CKM matrix. We envisage a scenario where charged weak interactions are $\mathrm{CP}$ conserving, while all $\mathrm{CP}$-violating phenomena stem from physics beyond the SM. Along with unitarity, the CKM matrix $V_{\mathrm{CKM}}$ is already constrained by experimental data; in particular strange particle and $B$-meson decays lead to the measurement of the moduli of the CKM matrix elements $|V_{us}|$, $|V_{cb}|$, and $|V_{ub}/V_{cb}|$. Since the new physics contribution to the above decays in the most plausible extensions of the standard theory cannot compete with the SM tree-level processes, the determination of these matrix elements is likely to hold, even in the pre\-sence of physics beyond the SM \cite{nirquinn}. Further constraints on the CKM matrix arise from the strength of $\mathrm{CP}$ violation in the kaon sector measured by $|\epsilon_K|$ and from the observation of $B_d^0$--$\bar{B}_d^0$ mixing, along with the improved experimental lower limit of $\Delta m_{B_s}$. Within the SM, the experimental results for $\Delta m_{B_d}$ and $\Delta m_{B_s}$ can be used to extract the values of $|V_{td}|$ and $|V_{ts}|$, while the measured value of $|\epsilon_K|$ directly constrains the size of the CKM phase. What follows is a simple method of checking whether or not present experimental data already exclude the possibility of having a real CKM matrix. Let us assume that $V_{\mathrm{CKM}}$ is real\bem{This case has also been discussed, for example, in Ref.~\cite{relpap}.} and that the observed $\mathrm{CP}$ violation in the kaon sector arises solely from physics beyond the SM. The experimental information on $|V_{us}|$, $|V_{cb}|$, and $|V_{ub}/V_{cb}|$, together with the unitarity of $V_{\mathrm{CKM}}$, allows one to deduce the values for $|V_{td}|$ and $|V_{ts}|$. Moreover, a comparison of these values with those extracted from $\Delta m_{B_d}$ and $\Delta m_{B_s}$ measurements enables us to ascertain the viability of a real CKM matrix. One may be tempted to interpret any incompatibility between the two sets of $|V_{td}|$ and $|V_{ts}|$ values obtained above as an indication that the CKM matrix cannot be real. However, the situation is more involved for the following reason. Within the SM, $\Delta m_{B_d}$ and $\Delta m_{B_s}$ are generated only at one-loop level by the $W$-mediated box diagrams, and as a result, physics beyond the SM can significantly contribute to $\Delta m_{B_d}$ and $\Delta m_{B_s}$. Indeed, in many extensions of the standard theory, the same new physics that gives rise to $\epsilon_K$ is likely to provide significant contributions to $\Delta m_{B_d}$ and $\Delta m_{B_s}$. In this paper, we first analyse the somewhat unnatural situation where new physics contributes only to $\epsilon_K$ and not to the mass difference in the $B$ system. We show that in this case the currently available data already disfavour a real CKM matrix, thus confirming previous results \cite{che:etal,parodi:etal}. We then investigate the more plausible situation of a real CKM matrix, with new physics generating $\epsilon_K$ and also contributing to $\Delta m_{B_d}$ and $\Delta m_{B_s}$. In this case, as we shall see below, present experimental data are compatible with the scenario of a real CKM matrix. Our paper is organized as follows.~In \sec{general}, we present a general analysis of $\Delta B=2$ transitions in the presence of new physics. Our results are model independent and apply to a wide class of models with a real CKM matrix. Section \ref{general:realCKM} deals with multi-Higgs-doublet models in which $\mathrm{CP}$ is softly broken. Assuming that $\mathrm{CP}$ is a symmetry of the Lagrangian, spontaneously broken by the vacuum, we discuss the possibility of having a real CKM matrix in a natural way. In \sec{2HDM}, we illustrate how the experimental constraints on $\epsilon_K$, $\Delta m_K$, and $\Delta m_{B_d}$ can be satisfied with a real CKM matrix in the context of a simple model with two Higgs doublets, and analyse the predictions for $\mathrm{CP}$ asymmetries in non-leptonic neutral $B$ decays that will be measured in upcoming $B$ experiments. Finally, we present our conclusions in \sec{summary}. \section{General analysis}\label{general} \subsection{Basic formulae and notation} We first carry out a general analysis of the new physics contribution to $\Delta B=2$ transitions (see, e.g., Refs.~\cite{babar,grossmanetal}). The off-diagonal element of the $B_q$ mass matrix can be written as\bem{For notational simplicity, we will omit the $q$ subscript for $H_{\mathrm{eff}}$.} \begin{equation}\label{mdef} M_{12}(B_q)\equiv M_{12}^{\mathrm{New}}(B_q) + M_{12}^{\mathrm{SM}}(B_q) =\frac{\braket{B_q}{H_{\mathrm{eff}}}{\bar{B}_q}}{2m_{B_q}}\ , \end{equation} where $q=d$ or $s$, and the effective Hamiltonian has the form \begin{equation}\label{fullheff} H_{\mathrm{eff}} = \heff^{\mathrm{SM}} + \heff^{\mathrm{New}}\ . \end{equation} It is customary to parametrize the new physics contribution appearing in \eq{mdef} through \begin{equation}\label{np} r_q^2 \exp(2i\theta_q) \equiv \frac{M_{12}(B_q)}{M_{12}^{\mathrm{SM}}(B_q)}\ , \end{equation} so that \begin{equation} \Delta m_{B_q}= 2|M_{12}^{\mathrm{SM}}|_q r^2_q\ . \end{equation} Note that $r_q^2< 1$ arises if the new physics amplitudes interfere destructively with those of the SM. The contribution of the SM to $B_q^0$--$\bar{B}_q^0$ mixing is given by \begin{equation}\label{m12sm} M_{12}^{\mathrm{SM}} (B_q)=\frac{G_F^2}{12\pi^2}\eta_B m_{B_q}m_W^2 f_{B_q}^2B_{B_q}S_0(x_t) (V_{tq}^{}V_{tb}^*)^2, \quad x_t = \ol{m}_t^2/m_W^2\ , \end{equation} where the various factors appearing in \eq{m12sm} may be found elsewhere \cite{qcd:factors,buras}. In fact, we have \begin{equation}\label{m12:bd} M_{12}^{\mathrm{SM}} (B_d)= 3.1\times 10^3\ \mathrm{ps}^{-1} \left[\frac{f_{B_d}\sqrt{B_{B_d}}}{200\ {\mathrm{MeV}}}\right]^2 \left[\frac{S_0(x_t)}{2.36}\right] \left[\frac{\eta_B}{0.55}\right] (V_{td}^{}V_{tb}^*)^2\ , \end{equation} and \begin{equation} M_{12}^{\mathrm{SM}} (B_s) = 4.5\times 10^3\ \mathrm{ps}^{-1} \left[\frac{f_{B_s}\sqrt{B_{B_s}}}{240\ {\mathrm{MeV}}}\right]^2 \left[\frac{S_0(x_t)}{2.36}\right] \left[\frac{\eta_B}{0.55}\right] (V_{ts}^{}V_{tb}^*)^2\ . \end{equation} Introducing the SU(3) breaking parameter \begin{equation} \xi_s\equiv \frac{f_{B_s}\sqrt{B_{B_s}}}{f_{B_d} \sqrt{B_{B_d}}}\ , \end{equation} we find the ratio \begin{equation}\label{rdrs} \frac{\Delta m_{B_s}}{\Delta m_{B_d}} =\xi^2_s\frac{m_{B_s}}{m_{B_d}} \left|\frac{V_{ts}}{V_{td}}\right|^2\frac{r_s^2}{r_d^2}\ . \end{equation} Finally, the $\mathrm{CP}$-violating asymmetry between $B^0$ and $\bar{B}^0$ mesons decaying to $\mathrm{CP}$ ei\-gen\-states $f_\mathrm{CP}$, such as $J/\psi K^0_S$ or $\pi^+\pi^-$, in the presence of new physics is given by \begin{equation}\label{def:asym} a_{f_\mathrm{CP}}(t)=\frac{\Gamma(B^0_d(t)\to f_\mathrm{CP})-\Gamma(\bar{B}^0_d(t)\to f_\mathrm{CP})}{\Gamma(B^0_d(t)\to f_\mathrm{CP})+\Gamma(\bar{B}^0_d(t)\to f_\mathrm{CP})}= -a_{f_\mathrm{CP}} \sin (\Delta m_{B_d}t)\ , \end{equation} with \begin{equation}\label{asym} a_{J/\psi K^0_S}= \sin 2(\beta +\theta_d+ \theta_K), \quad a_{\pi^+\pi^-} = \sin 2(\alpha -\theta_d)\ . \end{equation} Here we have ignored penguin contributions as well as final-state interaction phases. In addition, we have assumed that the direct decay amplitudes are dominated by the SM tree diagrams. The phase $\theta_K$ in the above expression is caused by the new physics contribution (normalized to the SM amplitude) to $K^0$--$\bar{K}^0$ mixing, i.e. the mixing parameter $(q/p)_K$ picks up another phase. However, given the experimental information on $\epsilon_K$, we expect the phase $\theta_K$ to be of ${\mathcal O}(10^{-3})$. We will return to this point in \sec{2HDM}. \subsection{New physics contribution and experimental data} As mentioned above, we will deal with a real CKM matrix. In this case, we have \begin{equation} \left|\frac{V_{ub}}{V_{cb}}\right|= \lambda |\rho|\ , \end{equation} and (neglecting higher order terms in $\lambda$) \begin{equation} V_{td}\simeq A\lambda^3\left[1-\rho\left(1-\frac{1}{2}\lambda^2\right)\right],\quad V_{ts}\simeq -A\lambda^2\left[1-\frac{1}{2}\lambda^2(1-2\rho)\right]\ , \end{equation} with the Wolfenstein parameters $A$, $\lambda$, and $\rho$ \cite{ckm:wolfenstein}. In our numerical analysis, we use the following input parameters \cite{babar,buras,pdg,alexander,sachrajda}: \begin{equation} \left|\frac{V_{ub}}{V_{cb}}\right| = \left|\frac{V_{ub}}{V_{cb}}\right|_{\mathrm{T}}\pm 0.005, \quad |V_{cb}| = 0.0395 \pm 0.0017\ , \end{equation} \begin{equation} A= 0.819 \pm 0.035, \quad \lambda = 0.2196\pm 0.0023, \quad \ol{m}_t= 167\pm 6\ {\mathrm{GeV}}\ , \end{equation} \begin{equation}\label{exp:epsi} |\epsilon_K| = (2.280\pm 0.019)\times 10^{-3}\ , \end{equation} \begin{equation}\label{exp:deltamk} \Delta m_K= (0.5301\pm 0.0014)\times 10^{-2}\ \mathrm{ps}^{-1}\ , \end{equation} \begin{equation}\label{bound:bs} \Delta m_{B_d}= 0.471\pm 0.016\ \mathrm{ps}^{-1},\quad \Delta m_{B_s}> 12.4\ \mathrm{ps}^{-1}\ (\cl{95})\ , \end{equation} and consider the ranges \begin{equation}\label{theo:var1} 0.06\leqslant \left|\frac{V_{ub}}{V_{cb}}\right|_{\mathrm{T}} \leqslant 0.10,\quad 160\ {\mathrm{MeV}}\leqslant f_{B_d}\sqrt{B_d}\leqslant 240\ {\mathrm{MeV}}\ , \end{equation} \begin{equation}\label{theo:var2} 1.12\leqslant \xi^2_s\leqslant 1.48\ . \end{equation} By using the experimentally measured values for $|V_{ub}/V_{cb}|$, $|V_{cb}|$, and $\Delta m_{B_d}$ (ignoring for the moment the constraint imposed by the $B_s$ system), we can perform a fit with two parameters $\rho$ and $r_d^2$. The results of the fit are shown in \fig{fig1}, \begin{figure} \centerline{\psfig{figure=figure1.ps,width=3.2in,angle=0}} \vspace{0.4cm} \centerline{\psfig{figure=figure2.ps,width=3.2in,angle=0}} \caption[]{The allowed region (at $\cl{95}$) in the $(\rho, r_d^2)$ plane for the Wolfenstein parameter $\rho>0$ (a) and $\rho<0$ (b), using the experimental values for $|V_{ub}/V_{cb}|$, $|V_{cb}|$, and $\Delta m_{B_d}$. Each contour corresponds to a variation of the theoretical input parameters, as described in the text [cf.~\eq{theo:var1}].\label{fig1}} \end{figure} from which we infer the upper and lower limits \begin{subequations}\label{bound:rho} \begin{equation} 0.24\lesssim \rho \lesssim 0.49, \quad 1.2 \lesssim r_d^2 \lesssim 6.5\ , \end{equation} \begin{equation} -0.49\lesssim \rho \lesssim -0.24, \quad 0.3\lesssim r_d^2 \lesssim 1.1\ . \end{equation} \end{subequations} Thus, for positive $\rho$ values, a new physics contribution to $\Delta m_{B_d}$ of at least $20\%$ of the SM contribution is required for consistency with current experimental data. Including the $\Delta m_{B_s}$ constraint, our analysis gives the lower bound as \begin{equation}\label{lower} \rho > 1-\sqrt{C_s\frac{r_s^2}{r_d^2}},\quad C_s = \frac{1}{\lambda^2}\frac{m_{B_s}}{m_{B_d}} \frac{\Delta m_{B_d}}{\Delta m_{B_s}}\,\xi_s^2\leqslant 0.8\, \xi_s^2\ . \end{equation} Two remarks are in order here. First, it is important to note that without new physics contributions to $\Delta m_{B_d}$ and $\Delta m_{B_s}$, i.e.~$r_d^2= r_s^2=1$, we obtain from \eq{lower} $\rho> - 0.08$. By using the bound of \eq{bound:rho}, one immediately sees that a real CKM matrix is disfavoured by present experimental data, as also reported in Refs.~\cite{che:etal,parodi:etal}. Second, for $r_s^2=1$ but $r_d^2\neq 1$, it follows from Eqs.~(\ref{bound:rho}) and (\ref{lower}) that either negative or positive values for $\rho$ are allowed by current experimental data. That is, a real CKM matrix is still consistent with the above measurements provided new physics contributes to both the $K$ and $B_d^0$ system. \section{A real CKM matrix}\label{general:realCKM} \subsection{Multi-Higgs-doublet models} This section is concerned with the question of how to construct models that naturally lead to a real CKM matrix, within the framework of multi-Higgs-doublet models (MHDMs).\bem{For a concise review of models with $\mathrm{CP}$ violation arising from the Higgs sector see, e.g., Ref.~\cite{bigi}.} The most general Yukawa couplings in the MHDM are given by the following Lagrangian (in the weak eigenbasis) \begin{equation}\label{yukawa:weak} {\mathcal L}_Y=-\sum_{j=1}^{M}\left(\bar{Q}_L\Gamma_j \phi_j n_R +\bar{Q}_L\tilde{\Gamma}_j \tilde{\phi}_j p_R\right) +\mathrm{H.c.}\ , \end{equation} with the left-handed quark doublet $Q_L = (p_L, n_L)^T$ and the right-handed quark singlets $p_R$ and $n_R$, where $p$ and $n$ denote the up- and down-type quarks respectively. $\Gamma_j$ and $\tilde{\Gamma}_j$ are matrices in flavour space, and the Higgs doublets are given by $\phi_j = (\phi_j^+, \phi_j^0)^T$, $\tilde{\phi}_j = i\sigma_2 \phi^*_j$, with $j=1, \dots, M$. In order to obtain a real CKM matrix in a natural way, it is particularly important to require $\mathrm{CP}$ invariance of the Lagrangian, only spontaneously broken by the vacuum. Without loss of generality, we may therefore assume that the above Yukawa matrices are real, while the vacuum expectation values (VEVs) of the neutral Higgs fields are given by $\langle{\phi_j^0}\rangle = v_j \exp(i\alpha_j)$, with $v^2=\sum_j v_j^2= (\sqrt{2}G_F)^{-1}$. Defining the Hermitian matrices \begin{eqnarray}\label{hu} H_u &\equiv& M_u M_u^{\dagger} = \frac{1}{2}\bigg\{\sum_i v_i^2\tilde{\Gamma}_i\tilde{\Gamma}_i^T + \sum_{j>i}v_i v_j \Big[\cos(\alpha_j-\alpha_i)(\tilde{\Gamma}_i\tilde{\Gamma}_j^T+ \tilde{\Gamma}_j\tilde{\Gamma}_i^T)\nonumber\\ &&\mbox{}\hspace{2.15cm}+i \sin(\alpha_j-\alpha_i) (\tilde{\Gamma}_i\tilde{\Gamma}_j^T-\tilde{\Gamma}_j\tilde{\Gamma}_i^T)\Big]\bigg\}\ , \end{eqnarray} and \begin{eqnarray}\label{hd} H_d &\equiv& M_d M_d^{\dagger} = \frac{1}{2} \bigg\{\sum_i v_i^2\Gamma_i\Gamma_i^T + \sum_{j>i}v_i v_j \Big[\cos(\alpha_j-\alpha_i)(\Gamma_i\Gamma_j^T+\Gamma_j\Gamma_i^T)\nonumber\\ &&\mbox{}\hspace{2.15cm}+i \sin(\alpha_j-\alpha_i)(\Gamma_j\Gamma_i^T- \Gamma_i\Gamma_j^T)\Big]\bigg\}\ , \end{eqnarray} where $M_u$ and $M_d$ are the non-diagonal up- and down-type mass matrices respectively, one can show that within the three-generation SM a necessary and sufficient condition for $\mathrm{CP}$ invariance in the charged gauge interactions is given by \cite{commutator} \begin{equation}\label{condition} T\equiv {\mathrm{Tr}}\,[H_u, H_d]^3=0\ . \end{equation} Note that this is independent of the number of Higgs doublets which generate masses for the fer\-mi\-ons. Hence, $T\neq 0$ implies $\mathrm{CP}$ violation mediated by charged gauge interactions, while $T=0$ guarantees a real CKM matrix. From \eq{condition} it is evident that the simplest and most natural way of obtaining a real CKM matrix (i.e. without any fine-tuning) is by introducing additional symmetries with the effect that \begin{equation}\label{relation:gamma} \Gamma_j\Gamma_i^T=\Gamma_i\Gamma_j^T \quad \mathrm{and}\quad \tilde{\Gamma}_i\tilde{\Gamma}_j^T =\tilde{\Gamma}_j\tilde{\Gamma}_i^T\ . \end{equation} Various multi-Higgs-doublet models where the condition in \eq{relation:gamma} is satisfied as a result of a discrete symmetry have been discussed in the literature. For example, the simplest models belonging to this class contain two Higgs doublets and a $Z_2$ symmetry \cite{luis} or three Higgs doublets and a $Z_3$ symmetry \cite{gustavo-sanda}. In these models, there are flavour-changing neutral currents and the vacuum leads to spontaneous $\mathrm{CP}$ violation, but the CKM matrix is real as a result of the discrete symmetry. \subsection{Supersymmetric standard model} The supersymmetric SM with spontaneous $\mathrm{CP}$ violation (SCPV) is another example where the CKM matrix is real provided that only two Higgs doublets are introduced. In the minimal supersymmetric standard model (MSSM), the tree-level vacua are all $\mathrm{CP}$ conserving so that SCPV can only occur if radiative corrections to the Higgs potential are taken into account. This inevitably leads to a mass of the lightest neutral Higgs boson which has already been ruled out by experiment \cite{susy:mssm}. On the other hand, in the next-to-minimal supersymmetric standard model (NMSSM) with one or more singlet fields in addition to the two doublets \cite{susy:nmssm}, SCPV can occur even at tree level. Whether or not this scenario is still consistent with present experimental bounds on the Higgs mass depends on the specific model, and we refer the interested reader to Ref.~\cite{disc:nmssm}. The important point to emphasize is that in the supersymmetric SM, provided there are only two Higgs doublets, but allowing for an arbitrary number of singlets, the CKM matrix is real, even if there is a physically meaningful relative phase between the VEVs of the two doublets. This can be readily verified by noting that the above phase can be eliminated from the quark mixing matrix \cite{gustavo}. \section{A specific model}\label{2HDM} \subsection{Two-Higgs-doublet model} For definiteness, we will consider a minimal extension of the SM with SCPV and a real CKM matrix as a result of a $Z_2$ symmetry \cite{luis}. Our main aim is to show that, within the framework of a specific two-Higgs-doublet model (2HDM), it is possible to have a real CKM matrix in a natural way, with $\epsilon_K$ and $\Delta m_{B_d}$ generated by physics beyond the SM, while taking into account the constraints discussed in \sec{general}. As we shall see below, $\mathrm{CP}$ violation arises solely from the exchange of neutral Higgs bosons, thereby inducing flavour-changing neutral current (FCNC) interactions at tree level. We start with the mass matrices $M_u$ and $M_d$ which can be diagonalized by means of a biunitary transformation, namely \begin{equation} p_L\to V_L^u u_L, \quad n_L\to V_L^d d_L, \quad p_R\to V_R^u u_R, \quad n_R\to V_R^d d_R\ , \end{equation} so that \begin{equation}\label{massmatrix} M_d^{\mathrm{diag}}= V_L^{d\dagger}M_d V_R^d, \quad M_u^{\mathrm{diag}}= V_L^{u\dagger}M_u V_R^u\ , \end{equation} where \begin{equation} M_u= \frac{1}{\sqrt{2}}(v_1 \tilde{\Gamma}_1 +v_2 e^{-i\alpha} \tilde{\Gamma}_2), \quad M_d= \frac{1}{\sqrt{2}}(v_1 \Gamma_1 +v_2 e^{i\alpha}\Gamma_2)\ . \end{equation} In order to obtain the physical Higgs fields, it is useful to parametrize $\phi_1$ and $\phi_2$ in \eq{yukawa:weak} as follows \begin{equation} \phi_1=\frac{1}{\sqrt{2}} \left(\begin{array}{c} \sqrt{2}\varphi_1^+ \\ v_1+ \rho_1+i \eta_1 \end{array}\right) , \quad \phi_2=\frac{e^{i\alpha}}{\sqrt{2}} \left(\begin{array}{c} \sqrt{2}\varphi_2^+ \\ v_2+ \rho_2+i \eta_2\end{array}\right) . \end{equation} The pseudo-Goldstone bosons $G^+$ and $G^0$ in our model can then be found by introducing new bases, i.e. \begin{equation} \left(\begin{array}{c}G^+\\H^+\end{array}\right)= O \left(\begin{array}{c}\varphi_1^+\\ \varphi_2^+\end{array}\right) , \quad \left(\begin{array}{c}G^0\\ I\end{array}\right)= O \left(\begin{array}{c}\eta_1\\ \eta_2\end{array}\right) , \quad \left(\begin{array}{c}H^0\\ R\end{array}\right)= O \left(\begin{array}{c}\rho_1\\ \rho_2\end{array}\right) , \end{equation} with \begin{equation} O= \frac{1}{v}\left(\begin{array}{lr}v_1 & v_2\\ v_2& -v_1\end{array}\right)\ . \end{equation} The mass matrix of the neutral Higgs fields is diagonalized through an orthogonal matrix $U$ relating $H^0$, $R$, and $I$ to the mass eigenstates $H_i$ ($i=1, 2, 3$) via the relations \begin{equation} H^0 = \sum_{i=1}^3 U_{1i}H_i, \quad R = \sum_{i=1}^3 U_{2i}H_i, \quad I = \sum_{i=1}^3 U_{3i}H_i\ . \end{equation} The Higgs-boson interactions with the quarks are then governed by the following Lagrangian \begin{eqnarray}\label{yukawa:mass} {\mathcal L}_Y&=& (2\sqrt{2}G_F)^{1/2} \Bigg\{\Big[\bar{u}\Big(\Gamma^{u\dagger} V_{\mathrm{CKM}} P_L-V_{\mathrm{CKM}}\Gamma^d P_R\Big)d\Big] H^+ + \mathrm{H.c.}\Bigg\}\nonumber\\ &+&(\sqrt{2}G_F)^{1/2}\sum_{i=1}^3 \Bigg\{-U_{1i}H_i\Big[\bar{d}M_d^{\mathrm{diag}}d+\bar{u}M_u^{\mathrm{diag}}u\Big]\nonumber\\ &&\hspace{1cm}\mbox{}-U_{2i}H_i\Big[\bar{d}\big(\Gamma^d P_R+ \Gamma^{d\dagger} P_L\big)d + \bar{u}\big(\Gamma^u P_R + \Gamma^{u\dagger} P_L\big)u\Big]\nonumber\\ &&\hspace{1cm}\mbox{}+ iU_{3i}H_i\Big[\bar{u}\big(\Gamma^u P_R - \Gamma^{u\dagger} P_L\big)u - \bar{d}\big(\Gamma^d P_R - \Gamma^{d\dagger} P_L\big)d\Big]\Bigg\}\ , \end{eqnarray} where $V_{\mathrm{CKM}}\equiv V_L^{u\dagger}V_L^d$ is the usual CKM matrix, $P_{L,R}=(1\mp\gamma_5)/2$, and \begin{eqnarray} \Gamma^u=V_L^{u\dagger}\left(\tilde{\Gamma}_1\frac{v_2}{\sqrt{2}} -e^{-i\alpha}\tilde{\Gamma}_2\frac{v_1}{\sqrt{2}}\right)V_R^u,\ \Gamma^d=V_L^{d\dagger}\left(\Gamma_1\frac{v_2}{\sqrt{2}} -e^{i\alpha}\Gamma_2\frac{v_1}{\sqrt{2}}\right)V_R^d\ .\nonumber\\ \end{eqnarray} In order for the above Yukawa coupling matrices to fulfil the relation of \eq{relation:gamma}, we impose an extra discrete $Z_2$ symmetry on the quark sector [see \eq{yukawa:weak}] under which $\phi_2\to -\phi_2$ and $n_{R3}\to -n_{R3}$ , while all other fields remain unchanged.\bem{An alternative assignment is to choose both $n_{R3}$ as well as $p_{R3}$ to be odd under $Z_2$. The advantage of this choice is that it leads to a naturally vanishing parameter $\bar{\theta}\equiv \theta_{\mathrm{QFD}} + \theta_{\mathrm{QCD}}$ at tree level, where $\theta_{\mathrm{QFD}}= \arg[\det(M_u M_d)]$.} Denoting the ratio of the VEVs of the two Higgs bosons by $\tan\beta\equiv v_2/v_1$, we find \begin{equation}\label{gammau:diag} \Gamma^u = M_u^{\mathrm{diag}}\tan\beta\ . \end{equation} Hence there are no FCNC interactions in the up-quark sector (i.e. no $\mathrm{CP}$ violation in $D^0$--$\bar{D}^0$ mixing). Turning to the down-quark sector, the CKM matrix can be made real\bem{The one-loop corrections to the parameter $J_{\mathrm{CP}}$ \cite{jarlskog} are equal to zero within the 2HDM in question \cite{luis:private}.} by redefining the phase of the quark field through $n_{R3}\to e^{-i \alpha}n_{R3}$ so that $V_L^{u,d}$ and $V_R^{u,d}$ are real and orthogonal matrices, while $\Gamma_d$ takes the explicit form \begin{equation}\label{gammad} \Gamma^d= \left(\begin{array}{ccc} m_d(\tan\beta-Sx^2) & -m_d S\beta_K & -m_d S\beta_d\\ -m_s S\beta_K & m_s(\tan\beta-Sy^2) & -m_s S\beta_s\\ -m_b S\beta_d & -m_b S\beta_s & m_b(\tan\beta-Sz^2) \end{array}\right) . \end{equation} Here $S=\tan\beta + \cot\beta$, with $S \geqslant 2$, and the $\beta_n$ $(n=K, d, s)$ are defined as \begin{equation}\label{beta:def} \beta_K= xy, \quad \beta_d= xz, \quad \beta_s=yz\ , \end{equation} where $x, y, z$ refer to the elements of the third column of $V_R^{d\dagger}$ with $x^2+y^2+z^2=1$. Consequently, the flavour-changing couplings of the three real neutral Higgs fields are constrained to obey \begin{equation}\label{beta:constraint} \beta_K\beta_d\beta_s = \left(\beta_K\beta_d\right)^2+\left(\beta_K\beta_s\right)^2 +\left(\beta_d\beta_s\right)^2, \quad |\beta_n|\leqslant 1/2\ . \end{equation} The crucial point here is that the mass matrix $M_d$, given in \eq{massmatrix}, is in general not a Hermitian matrix and consequently $V_R^d\neq V_L^d$, which in turn implies that $V_R^d$ is unrelated to $V_{\mathrm{CKM}}$. As a result, the $\beta_n$ above can be taken as free parameters, constrained only by condition (\ref{beta:constraint}). This is an important feature because the strength of the neutral Higgs contributions to $\Delta m_K$ and $\Delta m_{B_{d, s}}$ (see below) are proportional to $\beta_K$ and $\beta_{d, s}$ respectively. In fact, the neutral Higgs exchange induces a new physics contribution, in addition to the SM effective $\Delta S = 2$ Hamiltonian in \eq{fullheff}, given by \begin{eqnarray}\label{szwei} \lefteqn{\heff^{\mathrm{New}} =-\frac{G_F}{\sqrt{2}}\,m_s^2 (\beta_K S)^2 \sum_{i = 1}^{3}\frac{1}{M_i^2}}&&\nonumber\\ &&\times\bigg\{ C_i^{LL}(\bar{d}P_L s)^2 + C_i^{RR}(\bar{d}P_Rs)^2 +C_i^{LR}(\bar{d}P_Ls)(\bar{d}P_R s)\bigg\} + \mathrm{H.c.}\ , \end{eqnarray} with \begin{eqnarray} C_i^{LL}= 2(U_{2i}-iU_{3i})^2,\ C_i^{RR}= 2\zeta^2(U_{2i}+iU_{3i})^2, \ C_i^{LR}= 4 \zeta (U_{2i}^2+U_{3i}^2)\ ,\nonumber\\ \end{eqnarray} where we have introduced the shorthand $\zeta = m_d/m_s$. The effective Ha\-mil\-to\-nian inducing $B_q^0$--$\bar{B}_q^0$ mixing can be obtained from the above by replacing \begin{eqnarray} \label{replacement} s\to b,\ m_s \to m_b,\ \beta_K\to \beta_d;\quad s\to b,\ m_s \to m_b,\ m_d \to m_s,\ \beta_K\to \beta_s\ ,\nonumber\\ \end{eqnarray} for the $B_d^0$ and $B_s^0$ system respectively. To estimate the order of magnitude of the hadronic matrix elements resulting from \eq{szwei}, we rely on the vacuum insertion approximation. Defining the $F$-meson decay constant $f_F$ ($F=K, B_q$) via \begin{equation} \braket{F^0}{\bar{q}\gamma_{\mu}\gamma_5q'}{0}\braket{0}{\bar{q}\gamma^{\mu}\gamma_5q'}{\bar{F}^0} = f_F^2 m_F^2, \quad q= d, s,\ q' = b, s,\ q'\neq q\ , \end{equation} and employing the equation of motion, we get \begin{equation} \braket{F^0}{\bar{q}\gamma_5q'}{0}\braket{0}{\bar{q}\gamma_5q'}{\bar{F}^0} = - \left(\frac{m_F}{m_q + m_{q'}}\right)^2 f_F^2 m_F^2 \ . \end{equation} Thus, we obtain for the relevant matrix elements of the four-quark operators \begin{equation} \braket{F^0}{(\bar{q}P_Lq')(\bar{q}P_Rq')}{\bar{F^0}}= \frac{1}{2}\left[\left(\frac{m_F}{m_q + m_{q'}}\right)^2 + \frac{1}{6}\right] f_F^2B_{1F} m_F^2\ , \end{equation} \begin{equation} \braket{F^0}{(\bar{q}P_{L,R}q')(\bar{q}P_{L,R}q')}{\bar{F^0}}= -\frac{5}{12}\left(\frac{m_F}{m_q + m_{q'}}\right)^2 f_F^2 B_{2F}m_F^2\ , \end{equation} where $B_{iF}$ are the bag parameters. In what follows we will use $B_{iF}=1$, as de\-ter\-mined by the vacuum insertion approximation, which is sufficient for our purposes. The new physics contribution to $M_{12}$ in the $K$ system, \eq{szwei}, then takes the form \begin{eqnarray} \lefteqn{\left.M_{12}^{\mathrm{New}}\right|_K = \frac{G_F}{\sqrt{2}}\, m_s^2(\beta_K S)^2f_K^2 m_K \sum_{i=1}^3 \frac{1}{M_i^2}}&&\nonumber\\ &\times& \Bigg\{\Bigg(\frac{5}{12}\Big[(U_{2i}-iU_{3i})^2+ \zeta^2(U_{2i}+iU_{3i})^2\Big] -\zeta(U_{2i}^2+U_{3i}^2)\Bigg) \Bigg(\frac{m_K}{m_s + m_d}\Bigg)^2\nonumber\\ &-&\frac{1}{6}\zeta (U_{2i}^2+U_{3i}^2)\Bigg\}\ , \end{eqnarray} which in the limit $\zeta \ll 1$ reduces to \begin{equation}\label{m12:Knew} \left.M_{12}^{\mathrm{New}}\right|_K\simeq \frac{5}{12}\,\frac{G_F}{\sqrt{2}}f_K^2 m_K^3 (\beta_K S)^2\sum_{i=1}^3\frac{1}{M_i^2}(U_{2i}^2-U_{3i}^2- 2iU_{2i}U_{3i})\ . \end{equation} As can be seen from Eqs.~(\ref{replacement}) and (\ref{m12:Knew}), the phase in the off-diagonal element $M_{12}$ is common to both the $K$ and $B$ system, and we may therefore write \begin{subequations}\label{m12:gen} \begin{equation} M_{12}^n = |M_{12}^{\mathrm{New}}|_n e^{i\delta}+ M_{12}^{\mathrm{SM}}|_n\equiv M_{12}^{\mathrm{SM}}|_n r^2_n e^{2i\theta_n}, \quad n= K, d, s\ , \end{equation} with \begin{equation} \theta_n = \frac{1}{2}\arctan\left(\frac{R_n\sin\delta}{1+R_n\cos\delta}\right), \quad r^2_n = \sqrt{1+ 2R_n\cos\delta+R^2_n}\ , \end{equation} \end{subequations} where $R_n\equiv |M_{12}^{\mathrm{New}}|_n/M_{12}^{\mathrm{SM}}|_n$. \subsection{Constraints from $\epsilon_K$, $\Delta m_K$, and $\Delta m_{B_d}$} We now proceed to discuss various constraints on the parameters of the 2HDM. The observed $\mathrm{CP}$ violation in $K^0$--$\bar{K}^0$ mixing is described by (for $\epsilon'\ll\epsilon_K$) \begin{equation}\label{formula:epsi} \epsilon_K\simeq\frac{e^{i\pi/4}}{\sqrt{2}} \left(\frac{{\mathrm{Im}}\, M_{12}^K}{\Delta m_K}\right)\ ,\end{equation} whereas the mass difference in the $K$ system has the form \begin{equation}\label{massdiff:kaons} \Delta m_K\simeq 2|{\mathrm{Re}}\, M_{12}^K|\ . \end{equation} Using the experimental information on $\epsilon_K$ and $\Delta m_K$, Eqs.~(\ref{exp:epsi}) and (\ref{exp:deltamk}), as well as the central value $f_K= 160 \ {\mathrm{MeV}}$, we obtain \begin{equation}\label{cpconstraint} (\beta_K S)^2\sum_{i=1}^3 \frac{U_{2i}U_{3i}}{(M_i/{\mathrm{TeV}})^2}\simeq 5.2 \times 10^{-4}\ , \end{equation} and \begin{equation}\label{mass:diffK} (\beta_K S)^2 \sum_{i=1}^3\frac{U_{2i}^2-U_{3i}^2}{(M_i/{\mathrm{TeV}})^2} \lesssim 0.16\ . \end{equation} A few remarks are in order here. First of all, it can be readily shown that \cite{luis} \begin{equation}\label{cotalpha} \sum_{i=1}^3\frac{U_{2i}U_{3i}}{M_i^2}\propto \cot\alpha\ . \end{equation} Second, an analysis of the Higgs potential shows that in the limit of an exact $Z_2$ symmetry, the vacuum does not violate $\mathrm{CP}$ and the minimum of the potential is at $\alpha=\pi/2$, so that $\cot\alpha =0$. Allowing only a soft breaking of the $Z_2$ symmetry by a dimension-two term in the Higgs potential, spontaneous $\mathrm{CP}$ violation can be generated \cite{luis,gi}. This in turn implies that one may have $\cot\alpha \ll 1$ in a natural way (in the sense of 't Hooft \cite{thooft}), since the limit $\cot\alpha =0$ corresponds to the restoration of an exact $Z_2$ symmetry of the Lagrangian. This is the mechanism proposed by Branco and Rebelo \cite{gi} in order to naturally suppress the strength of $\mathrm{CP}$ violation. Therefore, the limit of \eq{cpconstraint}, which is due to the $\epsilon_K$ measurement, does not necessarily imply very large values of the Higgs mass. On the other hand, we do not consider any accidental cancellation between the elements $U_{2i}$ and $U_{3i}$ appearing in \eq{mass:diffK} because it would require additional fine-tuning of the parameters in the Higgs potential. Furthermore, we should emphasize here that there are also contributions from box diagrams with $W$-boson and non-SM charged Higgs particle exchange, as well as long-distance contributions, which affect only the \emph{real part} of $M_{12}^K$. Thus the most stringent limit on $M_i$ comes from the experimentally measured value of $\Delta m_K$, resulting in the upper bound of \eq{mass:diffK}. So, taking $(\beta_K S)$ to be of order unity, the Higgs masses have to satisfy $M_i \gtrsim \mathrm{few}\ {\mathrm{TeV}}$, whereas if we assume $(\beta_K S) \sim {\mathcal O}(10^{-1})$, then it is sufficient to demand that $M_i\gtrsim 250\ {\mathrm{GeV}}$. Turning to the $B$ system, the mass difference is given by \begin{equation}\label{massdiff:b} \Delta m_{B_q}= 2|M_{12}^q|,\quad q= d, s\ , \end{equation} which is valid for $\Gamma_{12}^q/M_{12}^q\ll 1$. Recall from \sec{general} that, in order for a model with a real CKM matrix to be consistent with present experimental data, a new physics contribution to $\Delta m_{B_d}$ equalling at least $20\%$ of the SM contribution is necessary when $\rho>0$. It can be shown that such a new contribution due to $M_{12}^{\mathrm{New}}|_d$ can be obtained in our model, while satisfying the bound of \eq{mass:diffK}. In fact, using Eqs.~(\ref{m12:bd}), (\ref{replacement}), (\ref{m12:Knew}), (\ref{m12:gen}), and (\ref{massdiff:b}), as well as central values for the various input parameters, one finds that even $(\beta_d/\beta_K)^2\sim {\mathcal O}(1)$ yields the necessary new physics contribution to $\Delta m_{B_d}$. Moreover, it is clear from the preceding discussion that there is a relation between $\theta_d$ and $\theta_K$ in the above 2HDM. Indeed, employing \eqs{m12:Knew}{formula:epsi}, and (\ref{massdiff:b}), we obtain \begin{equation}\label{Bsys:sintheta} \sin 2\theta_d = 2\sqrt{2}|\epsilon_K|\frac{\Delta m_K}{\Delta m_{B_d}} \left(\frac{m_{B_d}}{m_K}\right)^3\left(\frac{f_{B_d}}{f_K}\right)^2 \left(\frac{\beta_d}{\beta_K}\right)^2\ , \end{equation} and taking $f_K= 160\ {\mathrm{GeV}}$, $f_{B_d}=200\ {\mathrm{MeV}}$, we estimate \begin{equation}\label{beta:bounds} \left(\frac{\beta_d}{\beta_K}\right)^2\leqslant 7.4\ . \end{equation} We see that, due to the enhancement factor of $(m_{B_d}/m_K)^3$ in \eq{Bsys:sintheta}, $\mathrm{CP}$ violation in $B_d^0$-meson decays is not necessarily small. Before proceeding to discuss the $\mathrm{CP}$ asymmetries, we should mention that a more complete analysis of $\mathrm{CP}$ violation and physics beyond the SM has to take into account the recent measurement of $\epsilon'/\epsilon_K$ \cite{KTeV}. However, as pointed out by Buchalla \emph{et al.}\ \cite{buchallaetal} in the context of a specific 2HDM, a thorough investigation of the various contributions to direct $\mathrm{CP}$ violation in kaon decays, including renormalization group effects, is more involved and beyond the scope of the present paper.\bem{See also the discussion in Ref.~\cite{luis}.} \subsection{CP asymmetries and the 2HDM} We now turn to a study of the $\mathrm{CP}$ asymmetries. Due to the fact that the new physics contribution has a complex phase, the $\mathrm{CP}$ asymmetries in different $B$-meson decay channels are also affected. For example, using \eq{Bsys:sintheta} and the experimental value of \eq{exp:epsi}, we obtain for the $B_d^0$ system \begin{eqnarray}\label{thetasubd} \sin 2\theta_d \simeq 0.13 \left(\frac{\beta_d}{\beta_K}\right)^2 \left(\frac{f_{B_d}}{200\ {\mathrm{MeV}}}\right)^2 \left(\frac{160\ {\mathrm{MeV}}}{f_K}\right)^2\ . \end{eqnarray} Recall from \eq{asym} that in the case of a real CKM matrix, with $\alpha=\pi$ and $\beta=\gamma=0$, the $\mathrm{CP}$ asymmetries take the form \begin{equation} a_{J/\psi K^0_S}= \sin 2(\theta_d+ \theta_K), \quad a_{\pi^+\pi^-} = \sin 2(\pi -\theta_d)\ , \end{equation} and assuming the limit $\theta_K\ll \theta_d$, one finds \begin{equation}\label{predcn:multi} a_{J/\psi K^0_S}= - a_{\pi^+\pi^-}\ . \end{equation} We should emphasize that this result is a special feature of multi-Higgs-doublet models in which the CKM phase is absent, provided the new physics contributions cannot compete with the $W$-mediated tree diagrams of the SM. Note that due to the arbitrariness of the ratio $\beta_d/\beta_K$ in \eq{thetasubd}, the $\mathrm{CP}$ asymmetries in this model can in principle vary from zero to one. This should be compared with the SM estimate which predicts $a_{J/\psi K^0_S}$ to lie within the range $0.5 \leqslant a_{J/\psi K^0_S}\leqslant 0.9$ \cite{parodi:etal,SM:UT}. Once the two asymmetries $a_{J/\psi K^0_S}$ and $a_{\pi^+\pi^-}$ are measured, \eq{predcn:multi} will provide a clear test of the class of models where $\mathrm{CP}$ violation arises exclusively from flavour-changing neutral Higgs exchange. \section{Discussion and conclusions}\label{summary} One of the fundamental open questions in particle physics concerns the origin of $\mathrm{CP}$ violation. In particular, it is crucial to verify whether charged weak interactions violate $\mathrm{CP}$ or not. We have analysed the viability of a real CKM matrix, considering our present knowledge of the CKM matrix, derived from strange and $B$-meson decays, the experimental value of $\Delta m_{B_d}$, and the improved bound on $\Delta m_{B_s}$. We have shown that if one assumes that physics beyond the SM contributes only to $\epsilon_K$, then the real CKM matrix is clearly disfavoured by the above-mentioned data. However, if one makes the more plausible assumption that new physics also contributes to $\Delta m_{B_d}$ (and $\Delta m_{B_s}$), a real CKM matrix is still in keeping with present data. As a matter of fact, in order to fit the currently available data, it is crucial to have a new contribution to $\Delta m_{B_d}$ corresponding to at least $20\%$ (for $\rho > 0$), while a new contribution to $\Delta m_{B_s}$ is not relevant at this stage. This is because only an experimental lower bound on the oscillation frequency in the $B_s^0$ system is available. We have presented a general analysis which is quite model independent and applicable to a large class of models. This includes the supersymmetric SM with spontaneous $\mathrm{CP}$ violation, which has a real CKM matrix provided that only two standard Higgs doublets are introduced, apart from an arbitrary number of Higgs singlets. It is worth mentioning that within the supersymmetric SM it is possible to have the required additional contribution to $\Delta m_{B_d}$ \cite{gustavo:etal}. Moreover, we have shown that the assumption of spontaneous $\mathrm{CP}$ violation plays an extremely important role when the naturalness of a real CKM matrix is studied. For illustration, we have investigated a specific two-Higgs-doublet model where $\mathrm{CP}$ is spontaneously broken and the CKM matrix is real as a result of a $Z_2$ symmetry. In fact, we have shown that one can generate $\epsilon_K$ with simultaneous contributions to $\Delta m_{B_{d,s}}$ and obtain a good fit to the data. We are eagerly awaiting the upcoming experiments at the various $B$ factory facilities which will provide further constraints on the standard mechanism of $\mathrm{CP}$ violation as well as on alternative scenarios of $\mathrm{CP}$ breaking. \begin{ack} The work of F.\,K. has been supported by the TMR Network of the EC under contract ERBFMRX-CT96-0090. \end{ack}
\section{Physics motivation} Long-lived neutral or charged massive particles appear in many extensions of the MSSM. There are two main scenarios which result in a long lifetime of some of the SUSY partners. In the models that predict a degenerate mass spectrum of SUSY particles, light SUSY partners might be stable or long-lived, since the phase space would not allow for strong decays modes. It is particularly interesting if the NLSP is long-lived, since the observation of such an NLSP might not be doable via standard decay signatures expected for short-lived SUSY particles. As an example, many MSSM extensions predict degenerate mass spectrum of the neutralinos, in which case the following radiative decay could be a dominant decay mode of the second-lightest neutralino: \begin{equation} \tilde{\chi^0_2} \to \tilde\chi^0_1 \gamma. \label{eq:delayed-neutralino} \end{equation} Being electromagnetic, this decay mode is suppressed, and for fine splitting between the masses of the two lightest neutralinos, the decay constant ould be small enough to result in a long-lived $\tilde\chi^0_2$. Among other models which allow for a long-lived NLSP, are the GMSB scenarios in which a neutral NLSP (usually a neutralino) radiatively decays into a gravitino LSP: \begin{equation} \tilde\chi^0_1 \to \gamma\tilde G, \label{eq:delayed-GMSB} \end{equation} or a charged NLSP (usually the right-handed stau, $\tilde\tau_R$) decays into a $\tau$ and a gravitino NLSP: \begin{equation} \tilde\tau_R \to \tau\tilde G \label{eq:delayed-GMSB-tau} \end{equation} In the GMSB scenarios the gravitino mass is given by~(see, e.g., \cite{Feng}): $$ M_{\tilde G} = \sqrt{\frac{8\pi}{3}}\frac{F}{M_P}, $$ where $F$ is the vacuum expectation vaue of the dynamical supersymmetry breaking, and $M_P$ is Planck mass. Since most of the GMSB models predict $F \sim 10^[14]$~GeV$^2$, and $M_P \sim 10^{19}$~GeV, the gravitino is expected to be very light. The fact that the gravitino interacts with matter only weakly, could make the decays (\ref{eq:delayed-GMSB}) and (\ref{eq:delayed-GMSB-tau}) very slow. For example, the $c\tau$ for the decay (\ref{eq:delayed-GMSB-tau}) is given by~\cite{Feng} : $$ c\tau \approx 10~\mbox{km} \times \langle\beta\gamma\rangle \times \left[\frac{\sqrt{F}}{10^7~\mbox{GeV}}\right]^4 \times \left[\frac{100~\mbox{GeV}}{m_{\tilde\tau_R}}\right]^5, $$ and could be very large. Generally speaking, in different SUSY models, the $c\tau$ of the charged or neutral NLSP can be from subatomic distances to many kilometers. \section{Detection of the Delayed Decays} \label{sec:delayed} Detection and identification of long-lived particles depend on the decay path and the charge of this particle. A typical collider detector, e.g. the upgraded D\O\ apparatus~\cite{D0Upgrade}, has an inner Silicon Microstrip Tracker (SMT), capable of identifying secondary vertices from long-lived particle decays. The silicon detector is surrounded by a less precise Central Fiber Tracker (CFT), which is enclosed in the calorimeter. The calorimeter is surrounded by the muon system. Typical outer radii of the vertex detector, central tracking detector, calorimeter, and the muon system are $\sim 10$, 100, 200, and 1000~cm, respectively. In the case of the charged long-lived particles, one can identify the secondary decay vertex in the silicon vertex detector for $\gamma c\tau$ between $\sim 0.1$~cm and $\sim 10$~cm. Since we expect SUSY particles to be heavy, in what follows we will assume $\gamma \approx 1$, i.e. $\gamma c \tau \approx c\tau$. Both the CDF and D\O\ experiments will be capable of identifying such secondary vertices and possibly trigger on them. For 10~cm $\lesssim c\tau \lesssim 100$~cm, the charged long-lived particle will predominantly decay inside the outer tracking volume. The resulting characteristic signature is a kink in the outer tracker and large $dE/dx$ in the silicon vertex tracker and the inner layers of the outer tracker, typical for a slow-moving charged particle. While triggering on the kinks in the outer tracker won't be possible in Run II, there is a good chance that such kinks can be found offline, if the event is accepted by one of the standard triggers. Since a massive slow moving particle still has a significant momentum, one would likely trigger on such events using a single high-$p_T$ track trigger, or a designated $dE/dx$ trigger. For 100~cm $\lesssim c\tau \lesssim 200$~cm, the charged long-lived particle will decay inside the calorimeter, giving a jet-like energy deposition inside it. Identification of these particles will therefore rely on the fact that such a jet has only one track pointing to it (similar to one-prong $\tau$-decays); moreover, this track will have high $dE/dx$. In the case of the D\O\ detector, an additional $dE/dx$ measurement in the preshower detector can be used to aid the identification of such particles both offline and at the trigger level. A single high-$p_T$ track trigger, a designated $dE/dx$ trigger, or a $\tau$-trigger could be used to trigger on such events. Finally, for $c\tau \gtrsim 200$~cm, the long-lived particle will look stable from the point of view of the detector, and its identification will rely on a high $dE/dx$ track in the silicon vertex detector, the outer tracker, the calorimeter, and the muon system (if the $c\tau$ exceeds it outer radius). An additional time-of-flight (TOF) information from a designated TOF system (CDF) or muon system scintillators (D\O) could be used to trigger on such events and identify them offline. To summarize, both the CDF and D\O\ detectors will have very good capabilities for searches for charged long-lived particles with lifetimes from a few tens of picoseconds to infinity. Identification of the slow-moving particles using $dE/dx$ techniques in the case of the D\O\ detector is discussed in more detail in Section~\ref{sec:HIT-D0}. The situation is much more complicated for a neutral weakly interacting particle. First of all, if such a particle decays outside of the calorimeter, it can not be detected, and will look like a missing $E_T$ in the event, i.e. like the LSP. It's unlikely that one could identify the presence of a long-lived NLSP in such events, but it still might be possible to identify these events as the non-SM ones. If events with such a signature are found in Run II, one could conceivably install a wall of scintillating detectors far away from the main detector volume and try to look for a photon from the radiative decay in this additional scintillator~\cite{Gunion}. For the purpose of this report, however, we will focus only on the case of $c\tau < 2$~m, which roughly coincides with the outer radius of the D\O\ calorimeter. The detection technique for such decays relies heavily on the fine longitudinal and transverse segmentation of the preshower detector and the calorimeter, which is an essential and unique feature of the D\O\ detector. The signature for the radiative decay of a long-lived particle (e.g., (\ref{eq:delayed-GMSB}) or (\ref{eq:delayed-neutralino})), is a production of a photon with a non-zero impact parameter. If there was a way to identify the point of photon origin, one could single out such a delayed radiative decay corresponding to a very distinct and low background topology. As is shown in Section~\ref{sec:pointing}, the D\O\ calorimeter information, combined with the preshower information, can be used to achieve a very precise determination of the photon impact parameter. This technique would allow to identify long-lived neutral particles with 10~cm $\lesssim c\tau \lesssim 100$~cm, in the case of the upgraded D\O\ detector. The detectable $c\tau$ range can be further extended by a factor of two by looking for the photons from the radiative decay in the D\O\ hadronic calorimeter. The signature for such photons would look like a ``hot cell,'' i.e. an isolated energy deposit in one or two hadronic calorimeter cells. The reason that the EM energy deposition in the calorimeter is so isolated is the fact that each hadronic calorimeter cell contains many radiation lengths and completely absorbs an EM shower. The main background for this signature is production of high-$p_T$ $K_S^0$-mesons which decay into a pair of $\pi^0$'s corresponding to the 4$\gamma$ final state. Since the $K_S^0$ is significantly boosted, these four photons are highly collimated and will be identified as a single EM shower in the calorimeter. For a typical $K^0_S$ momentum of 25 GeV, which corresponds to $\sim 20$~GeV $E_T$ of the resulting $4\gamma$ system, the $\gamma c\tau$ is about 130~cm, i.e. most of the $K^0_S \to 4\gamma$ decays will occur in the hadronic calorimeter. This background, however, can be well predicted and also has a well defined $r$-dependence, where $r$ is the radial distance of the ``hot cell'' from the detector center. We therefore expect this background to be under control in Run II. Thus, the upgraded D\O\ detector will have unique capabilities for searching for neutral long-lived particles with 10~cm $\lesssim c\tau \lesssim 2$~m. Both the CDF and D\O\ could also explore the 0.1~cm $\lesssim c\tau \lesssim 10$~cm range by looking for a conversion of the photon from a radiative decay in the silicon vertex detector. If the photon converts, one can determine its direction and the impact parameter fairly precisely by looking at the tracks from the $e^+ e^-$-pair. One, however, would pay a significant price for being able to explore this region of $c\tau$, since the probability of photon conversion in the silicon vertex detector is only a few per cent. CDF could in principle use the conversion technique to explore higher values of the $c\tau$ by looking for conversions in the outer tracking chamber. However, due to the low conversion probability the sensitivity of the CDF detector in this region is by far lower than that of the D\O\ detector. Apart from being crucial for the study of physics with delayed radiative decays, the D\O\ detector ability for photon pointing is very attractive from other points of view. In the high luminosity collider environment the average number of interactions per crossing exceeds one. (It can be as high as five, for the 396~ns bunch spacing expected at the beginning of Run 2.) Therefore, each event will generally have several primary vertices with only one being from the high-$p_T$ interaction of physics interest, and the others being due to minimum bias $p\bar p$ collisions. The presence of multiple vertices creates a problem in choosing the right one for determination of the transverse energies of the objects. Photon pointing can solve this problem. It is especially important at the trigger level when the information about all the objects in the events is not generally available, and therefore high-$p_T$ objects which produce tracks in the tracking chambers can not always be used to pinpoint the hard scattering vertex. In some cases, for example the $\gamma + \met$ final state, there are no objects with tracks at all, so there is no way to determine what vertex the photon originated from without utilizing the calorimeter-based pointing. Not knowing which vertex is the one from high-$p_T$ collision results in the object $E_T$ mismeasurement, which is especially problematic for missing transverse energy calculations. Indeed, \met\ calculations rely heavily on the vertex position and picking the wrong one may result in significant missing transverse energy calculated in the event which in fact does not have any physics sources of real \met. Not does only this affect physics analyses for topologies with \met, especially the $\gamma + \met$ one, but it also results in a worsening of the \met\ resolution, hence a slower turn-on of the \met-triggers and ergo higher trigger rates. It is, therefore, very important to have a way of telling the high-$p_T$ primary vertex, and photon pointing is the only way of doing this in the $\gamma + \met$ case. Subsequent sections contain technical detail on the high-$dE/dx$ and delayed photon identification. \section{Photon Pointing at D\O} \label{sec:pointing} The importance of photon pointing was appreciated in some Run I analyses, particularly the studies of $Z(\nu\nu)\gamma$~\cite{Zg} and $\gamma\gamma$~\cite{monopoles}. We have utilized the fine longitudinal segmentation of the D\O\ EM calorimeter (which has four longitudinal layers) by calculating the c.o.g. of the shower in all four layers independently and then fitting the four spatial points to a straight line in order to determine the impact parameter and the $z$-position of the photon point-of-origin~\cite{EMVTX} (see Fig.~\ref{fig:emvtx}). The algorithm used for the c.o.g. finding is based on a logarithmic weighting of the energy deposition in the EM calorimeter cells which belong to the EM shower. The spatial resolution of the c.o.g. finding algorithm in four calorimeter layers averaged over central (CC) and forward (EC) rapidity range, as well as the geometrical parameters of the calorimeter layers are given in Table~\ref{table:EMVTX}. This study resulted in an algorithm, {\tt EMVTX}~\cite{EMVTX}, that has been used in several D\O\ analyses involving photons~\cite{Zg,monopoles}. \begin{figure}[thb] \vspace*{0.1in} \centerline{\protect\psfig{figure=emvtx.eps,width=\textwidth}} \caption{Side-view of the D\O\ calorimeter illustrating the application of the {\tt EMVTX} algorithm.} \label{fig:emvtx} \end{figure} \begin{table}[htb!] \begin{center} \caption{Geometry and average resolutions in the preshower detectors and the EM calorimeter layers.} \begin{tabular}{||l||l|l|l|l|l||} \hline Quantity & Preshower & EM1 & EM2 & EM3 & EM4 \\ \hline\hline \multicolumn{6}{||c||}{Central Region}\\ \hline $\langle R \rangle$ c.o.g. & 73.0~cm & 85.5~cm & 87.4~cm & 91.8~cm & 99.6~cm \\ $\sigma_z$ & 2.5~mm & 20~mm & 20~mm & 7.0~mm & 15~mm \\ $\sigma_{r\phi}$ & 1.5~mm & 17~mm & 17~mm & 3.5~mm & 7.5~mm \\ \hline Quantity & CPS & EM1 & EM2 & EM3 & EM4 \\ \hline \multicolumn{6}{||c||}{Forward Region}\\ \hline $\langle |Z| \rangle$ c.o.g. & 142.0~cm & 171.7~cm & 174.2~cm & 179.2~cm & 189.7~cm \\ $\sigma_r$ & 1.5~mm & 8.0~mm & 8.0~mm & 1.5~mm & 3.5~mm \\ $\sigma_{r\phi}$ & 2.5~mm & 7.0~mm & 7.0~mm & 1.0~mm & 2.8~mm \\ \end{tabular} \end{center} \label{table:EMVTX} \end{table} In order to study the improvement of photon pointing made possible by the utilization of the fine spatial resolution of the preshower detector, we have written a toy Monte Carlo (MC) simulation package which takes into account detector geometry, position error in calorimeter and preshower, as well as the primary vertex distribution. First, we compare the resolution obtained from the toy MC with the actual distributions obtained from $W \to e\nu$ events collected in Run 1. For electrons from $W$-events it is possible to determine the point of origin by using track information, which is quite precise. The difference between the $z$ position of the vertex obtained from tracking and from electron pointing as well as the signed impact parameter for the electron obtained by pointing are shown in Fig.~\ref{fig:emvtx_w}, for central electrons. (The positive sign corresponds to the impact parameter to the right of the center of the detector observed from the EM cluster location.) The distributions are fitted with Gaussian functions with the widths of $\sigma_z = 14.0$~cm and $\sigma_r = 9.5$~cm, as determined from toy MC. The data agrees well with the MC predictions and also appears Gaussian. An analogous comparison for forward electrons is shown in Fig.~\ref{fig:emvtx_w_ec}. The corresponding resolutions are $\sigma_z = 17.0$~cm and $\sigma_r = 4.5$~cm. The resolution on impact parameter improves in the forward region because the physical size of the calorimeter cells becomes smaller with an increase in $|\eta|$. In the $z$-direction, this effect is compensated by the small angle of the cluster pointing, which dilutes the $z$-resolution. \begin{figure}[hbt] \vspace*{-0.5in} \centerline{\protect\psfig{figure=emvtx_w.eps,width=4.0in}} \vspace*{-0.2in} \caption{Comparison of (a) the error on the $z$-position of the vertex, and (b) the signed impact parameter from Run 1 $W \to e\nu$ data, with the results of toy simulations, for central electrons.} \label{fig:emvtx_w} \end{figure} \begin{figure}[hbt] \vspace*{-0.5in} \centerline{\protect\psfig{figure=emvtx_w_ec.eps,width=4.0in}} \vspace*{-0.2in} \caption{Comparison of (a) the error on the $z$-position of the vertex, and (b) the signed impact parameter from Run 1 $W \to e\nu$ data, with the results of toy simulations, for forward electrons.} \label{fig:emvtx_w_ec} \end{figure} \begin{figure}[hbt] \centerline{\protect\psfig{figure=emvtx2.eps,width=\textwidth}} \caption{Side-view of the D\O\ calorimeter illustrating how the {\tt EMVTX} algorithm works.} \label{fig:emvtx2} \end{figure} \begin{figure}[hbt] \vspace*{-0.5in} \centerline{\protect\psfig{figure=emvtx_run2.eps,width=4.0in}} \vspace*{-0.2in} \caption{Run 2 simulation of (a) the error on the $z$-position of the vertex, and (b) the signed impact parameter, for central photons.} \label{fig:emvtx_run2} \end{figure} \begin{figure}[hbt] \vspace*{-0.5in} \centerline{\protect\psfig{figure=emvtx_run2_ec.eps,width=4.0in}} \vspace*{-0.2in} \caption{Run 2 simulation of (a) the error on the $z$-position of the vertex, and (b) the signed impact parameter, for forward photons. The slightly non-Gaussian shape is due to the change in the pointing resolution as a function of the photon rapidity.} \label{fig:emvtx_run2_ec} \end{figure} The central and forward preshower detectors of the D\O\ Upgrade provide a precision measurement of the photon cluster position. The resolution for a typical EM shower perpendicular to the preshower strip is $\sim 1$~mm. Taking into account the crossing angle between the $u$- and $v$-planes in the preshower detectors, the following resolutions in $r\phi$ and $z$ ($r$) can be obtained: $\sigma_{r\phi} = 1.5$~mm, $\sigma_z = 2.5$~mm (CPS) and $\sigma_{r\phi} = 2.5$~mm, $\sigma_r = 2.5$~mm (CPS). As one can see, the preshower cluster position measurement is superior to that obtained from the EM calorimeter; its position is also spatially separated from that in the calorimeter, as seen in Fig.~\ref{fig:emvtx2}. With the additional position measurement coming from the preshower detectors, the following pointing resolutions can be obtained for central and forward photons in Run 2: $\sigma_z = 2.2$~cm, $\sigma_r = 1.4$~cm (CC, see Fig.~\ref{fig:emvtx_run2}) and $\sigma_z = 2.8$~cm, $\sigma_r = 1.2$~cm (EC, see Fig.~\ref{fig:emvtx_run2_ec}). A very significant improvement in photon pointing (by a factor of six) is achieved by utilizing the additional position measurement provided by preshower detectors. The implementation of the photon-pointing algorithm can be done as early as at the trigger Level 3. An approximate algorithm that uses only the c.o.g. of the EM shower in the preshower and in the third layer of the EM calorimeter could be used to decrease the amount of calculations and, hence, the decision-making time. Monte Carlo simulations show that the impact parameter and $z$-resolutions for a simplified algorithm are only 10\% worse than those obtained by complete five-point fit, which is quite satisfactory for trigger purposes. Having the precise vertex information from photon-pointing for photon triggers at Level 3, we will also recalculate the \met based on this vertex, and that would significantly improve the turn-on of the \met part of less inclusive triggers which would require an EM cluster and \met. \section{Detection of Slow-Moving Massive Charged Particles in D\O} As described in Section~\ref{sec:delayed}, long-lived charged particles have a variety of characteristics which enable their identification, particularly in the analysis stage when the event's complete data set is available. Depending on the lifetime, a combination of time-of-flight, ionization ($dE/dx$) in several different detectors, and the muon-like penetration of a high momentum, isolated track all can be employed, with the additional presence of a kink where there is a decay within the detector volume. Even though discovery at the analysis stage seems possible, triggering is really crucial, if heavy stable particles produced at the Tevatron are to be detected. In this section we discuss the possibilities for triggering, and in particular, the use of $dE/dx$ in the hardware trigger as a tool for detecting these objects. The D\O\ trigger system is hierachial, with 3 levels, each level passing a small subset of the events it examines to the next level for further analysis. Thus Level 1 has a high input rate and examines a limited amount of information in making its decision, while subsequent levels have progressively lower input rates and spend longer analyzing more data. At Level 3, all the data digitized for the event is available and the selection is made running software algorithms written in high level code and derived from the offline analysis. We expect that the massive stable particles will be sufficiently rare, and their characteristics clear, such that separating candidates at Level 3 will be straight forward. However, we need to understand how to indentify these events in the hardware triggers so that they survive to Level 3. For Level 1 several tools are available to detect a massive stable particle. The Central Fiber Tracker provides track candidates binned in momentum to which it can apply an isolation criteria~\cite{Yuri}. The CFT has 80 trigger segments independently processed by the trigger. For a heavy stable particle one can select a high $p_T$ track with no other tracks in the home and adjacent segments, imposing an isolation of $\pm6.75^\circ$ in azimuthal angle~\cite{Yuri}. Association of a muon with such a track would provide an efficient Level 1 trigger for massive stable particles, produced centrally and in isolation. An additional tool~\cite{KJ} at Level 1 is the measurement of the time of flight (TOF) from scintillators associated with the muon detector and used primarily to reduce background in the large area muon detector from cosmic ray and beam associated accidentals. These counters cover much of the area just outside the calorimeter, inside the first (``A") layer of muon detectors, and completely outside the detector, on top of the muon ``C" layer planes. Electronics associated with the scintillation counters provides both a trigger gate and a TOF gate, thus giving time windows relative to the interaction time. Scintillator hits received within the TOF gate will have the time of flight measurement digitized and saved with the data, assuming the event otherwise passes Level 1. Typically the TOF gate will be set sufficiently wide to accept slow moving particles, namely at least 100 nsec. To contribute to the Level 1 trigger, however, the scintillator hit must occur within the trigger gate, which is necessarily much shorter (of order 25 nsec) because of the high rate of accidentals, particularly in the ``A'' layer counters. It may be possible to run with a considerable wider trigger gate for the ``C'' layer counters, given the lower expected accidental rate in these counters. In this case, TOF from the muon system would provide a useful tool for Level 1 triggering on slow particles. Given a Level 1 trigger, generated either through TOF or through an isolated stiff CFT track, the TOF data will be a useful tool for the Level 2 and Level 3 triggers. Beyond Level 1, the new Silicon Microstrip Tracker provides interesting possibilities for triggering on slow moving particles. Since energy loss of a slow moving particle drops as $1/\beta^2$ as a function of its velocity, the excellent $dE/dx$ energy resolution of the silicon chip provides a good handle for offline identification. For example, the energy deposited in one silicon layer is shown in Fig.~\ref{fig:MSP-ADC}, as measured in a test beam~\cite{Maria}. More importantly for the detection of slow moving particles, there are possibilities to exploit these measurements at the trigger level, with recent approval of the D\O\ Silicon Tracker Trigger (STT)~\cite{STT} as a component of the Level 2 trigger system. The hardware design for the STT may allow for the inclusion of several additional backplane lines to carry $dE/dx$ information along with the other data associated with each cluster of hits~\cite{Uli}. We have studied the basic capabilities of such trigger hardware to explore its potential for slow particle identification. \begin{figure}[hbt] \vspace*{0.1in} \centerline{\protect\psfig{figure=hit1.eps,width=\textwidth}} \caption{Energy deposition in one layer of the Silicon Microstrip Tracker in ADC counts, as measured in the test beam.} \label{fig:MSP-ADC} \end{figure} Our simulation of the STT slow particle trigger is based on a simple Monte Carlo generator which returns an ADC value appropriate for the spectrum shown in Fig.~\ref{fig:MSP-ADC}. We assume that two backplane lines per SMT hit are available, and we encode each ADC value into these two bits of data, or four bins. After trying various values for the three bin edges, we have chosen to define the bins by those ADC values below which lie 67\%, 95\%, and 99\% of the data. These cutoffs correspond to 38, 75, and 105 ADC counts (see Fig.~\ref{fig:MSP-ADC}). The STT Level 2 trigger processor finds tracks using four layers of silicon; so, in our model, the hardware would provide four samples of this two bit $dE/dx$ data for each track. At this point the trigger would use some algorithm to combine the four samplings most advantageously. The major concern for the correct identification of a slow moving particle is the likelihood of false signals from a minimum ionizing particle (MIP), due to an occasional response in the very long tail of the Landau energy loss distribution. Based on a few studies, our preliminary suggestion is simply to sum the three lowest values, rejecting the largest ADC count of the four. The data then would provide a parameter, related to a $dE/dx$ of the particle in the silicon tracker, which we call ``slowness," and which has 10 possible values (0..9). From our simulation we derive the distribution in slowness for a $\beta=1$ particle (equivalent to the test beam pion), as shown in Fig.~\ref{fig:MSP-slowness}. \begin{figure}[hbt] \vspace*{0.1in} \centerline{\protect\psfig{figure=hit2.eps,width=\textwidth}} \caption{Distribution of slowness for massive particles with $\beta = 0.4$ and 0.7, and test beam particles with $\beta = 1.0.$} \label{fig:MSP-slowness} \end{figure} The ADC response to the passage of a slow moving particle will be similar to the ADC distribution of Fig.~\ref{fig:MSP-ADC} scaled by the factor $1/\beta^2$. However, because of the very different kinematics (the slow moving particle is massive, typically 150 GeV) the energy loss distribution will have a less pronounced Landau tail. We make a very conservative assumption and use only the Gaussian component in generating ADC values for massive slow moving particles. Several resulting distributions in our $dE/dx$ trigger parameter, ``slowness,'' for particles with $\beta=0.4$ and $\beta=0.7$, are included in Fig.~\ref{fig:MSP-slowness}. There is a clear separation in this parameter compared to the $\beta=1$ distribution. To estimate the effectiveness of the STT $dE/dx$ trigger we consider a selection which tags as a slow particle those whose ``slowness" is greater than or equal to some value. We vary this selection to study the efficiency for slow particles (at highest possible $\beta$) while maintaining a strong rejection against $\beta=1$ particles. Figure~\ref{fig:MSP-eff} shows the acceptance as a function of a particle's velocity, for events with slowness $>4$. The appearance of a few $\beta=1$ particles but not other high $\beta$ events reflects the conservative use, for massive particles, of a Gaussian ADC response, rather than the Landau distribution, which is used to generate the ADC values from a MIP. \begin{figure}[hbt] \vspace*{0.1in} \centerline{\protect\psfig{figure=hit3.eps,width=\textwidth}} \caption{Efficiency as a function of particle $\beta$ for events with slowness $>3$. The open histogram corresponds to the case without angular smearing; the filled area shows the change when angular smearing of $\pm 45^\circ$ is taken into account.} \label{fig:MSP-eff} \end{figure} \begin{figure}[hbt] \vspace*{0.1in} \centerline{\protect\psfig{figure=hit4.eps,width=\textwidth}} \caption{Efficiency as a function of particle $\beta$ for the events with redefined slowness a) $>7$; b) $>8$. Angular smearing of $\pm 45^\circ$ is taken into account.} \label{fig:MSP-eff1} \end{figure} Despite the intent to trigger only on centrally produced objects, tracks will tend to be inclined to the silicon, increasing and broadening the $dE/dx$ response. In fact, the STT hardware will search for track candidates from adjacent barrels; so that tracks may be inclined to the normal by as much as $45^\circ$ although the STT will have no information about this angle. We have modeled the effect of inclined tracks by scaling the appropriate ADC response by $1/\cos\theta$, where $\cos\theta$ is generated uniformly between .707 and 1.0, or by $\eta$, where $\eta$ is generated uniformly between 0 and 0.88. The differences between smearing based on $\cos\theta$ and that based on $\eta$ are small. We present here results using $\cos\theta$ smearing for the centrally produced massive stable particles and $\eta$ smearing for the $\beta=1$ background. As seen in Fig.~\ref{fig:MSP-eff}, the effect of the variation in ADC response due to track inclination is only a small reduction in the rejection for $\beta=1$ particles; moreover, this effect helpfully raises the cutoff in $\beta$ for massive particles. Overall, this smearing does not appear to affect the STT ``slowness" trigger significantly. Including track inclination, the study suggests that the STT Level 2 $dE/dx$ trigger would provide a fully efficient tag for slow particles up to $\beta=0.7$ with an acceptance of $\beta=1$ particles less than $2\times 10^{-3}$. Because of its excellent rejection for $\beta=1$ particles, the Level 2 $dE/dx$ trigger will be a good means to select slow moving particles, independent of other criteria such as TOF or track isolation. However, if the particle is sufficiently long-lived to traverse the entire detector, it may be possible to relax the $dE/dx$ requirement. We have studied the effect of modifying the hardware ADC sampling, to explore widening the acceptance of heavy particles in $\beta$ at the expense of $\beta=1$ rejection. There is good physics motivation in doing so, as some GMSB models predict the production of heavy stable particles with $\beta$ in the range 0.8-0.9~\cite{Jianming}. Using ADC bins with edges corresponding to 50\%, 68\%, and 90\% of the distribution, we find good acceptance for massive particles at high $\beta$, as shown in Fig.~\ref{fig:MSP-eff1}, with rejection factors between 20 and 100, for $\beta=1$ particles. The above study illustrates the potential of a STT $dE/dx$ trigger. It does assume that the ADC distribution for a MIP is as seen in the test beam, and that differences in the silicon chip response over the detector won't significantly affect this distribution. The study suggests that track inclination may not be a serious problem. Futher, the ``slowness" flag derived with the STT from $dE/dx$ can in Level 2 be combined with other information (as of a straight, non-oblique and isolated muon) to provide a global Level 2 trigger. In summary, it seems promising that the STT could provide very useful $dE/dx$ information early in a slow particle's lifetime, which can be combined with other data to create an efficient trigger for these interesting objects. \label{sec:HIT-D0}
\section{Introduction} It is well known that a given physical theory describes well only a limited class of phenomena measured in certain range of precision and desagrees with other phenomena beyond this range which need a more general theory for their description. In general, the latter involves new fondamental constant which can be viewed as a kind of deformation parameter. The old theory can be recovered in the limit where the new fundamental constant vanishes ( which in general means that we do not reach enough precision to detect the effects tied to this fundamental constant). The special relativity and the quantum mechanics provide most famous examples of such a theory generalization. The special relativity may be regarded as a deformed theory of Galilei's relativity where the deformation parameter is the inverse of the velocity of light and the quantum mechanics is a deformed theory of the classical mechanics with the Planck constant as deformation parameter.\\ On the other hand, in view of the enormous and unsuccessful effort to overcome the divergence difficulties in the relativistic quantum field theories, it is believed that the framework of special relativity has to be changed.\\ In the past few years, attention has been paid to formulate the particle evolution in quantum Minkowski space through the construction of the q-analogs of the relativistic plane waves [1] or the Hilbert space representation of a q-deformed Minkowski space [2]. Despite all the theoretical interests, the relevance of the noncommutative Minkowski space-time and its transformations under the quantum Lorentz group to measurable effects in particle physics has not been discussed very much. \\ The above considerations make especially interesting the study of the noncommutative special relativity in the frame of the quantum Minkowski space-time and its transformations under the quantum Lorentz group to derive measurable observables describing the evolution of particles in the noncommutative space-time.\\ In this paper, we adapt the quantum mechanics axioms to the quantum coordinates of Minkowski space-time and their transformations under quantum Lorentz group to show how one can derive the quantum physical observables corresponding to those of the usual special relativity.\\ This paper is organized as follows: In section 2, we recall the basic properties derived from the correspondence between quantum $SL(2,C)$ and Lorentz groups developped in [3]. In section 3, we consider Hilbert spaces on which the generators of the Hopf algebras corresponding to the quantum Lorentz symmetries and the quantum Minkowski space-time act. We establish the transformation rules of states according to the properties of coordinate transformations which are given in terms of tensorial products. The structure of the commutation rules between the generators of the Lorentz symmetries and the coordinates of the Minkowski space-time permits us to introduce the velocity operator and to study the boost from which we establish the quantum analog of the lifetime dilatation formula of unstable particles. We end this section by nothing that this construction is also valid if we replace the coosrdinates $X_{N}$ of the quantum Minkowski space-time by the energy-momentum four-vector $P_{N}$. In this case we obtain the analog of the usual relativistic formulas giving the energy and the vector-momentum in terms of the mass and the velocity. In section 4, we carry on the investigation of the Hilbert space states to describe the evolution of particles in the quantum Minkowski space-time. From this state investigation, we establish the principle of the causality in the noncommutative special relativity and show that for a free particle moving in the noncommutative space-time, only the length of the velocity and one of its components can be measured exactly. In addition these observables present discret spectrums which lead to a quantized lifetime of unstable paricles. We conclude this paper in section 5 by discussing the case of particles moving in the light-cone. We also derive directly from this formalism construction of the quantum Lorentz subgroup of the three dimensional rotation of the space coordinates leaving the time coordinate invariant. This permits us to deduce the properties of quantum spheres. \section{The quantum Lorentz group} To explore the different properties of the quantum Lorentz group and the quantum Minkowski space time, it is very convenient to represent the generators $\Lambda_{N}^{~M}~(N,~M=~0,~1,~2,~3)$ of the quantum Lorentz group in terms of generators $M_{\alpha}^{~\beta}~(\alpha, \beta=1,2)$ and $(M_{\alpha}^{~\beta})^{\star} =M_{\dot{\alpha}}^{~\dot{\beta}}$ of the quantum $SL(2,C)$ group which is well known. Let us start by recalling in this section, some basic properties derived from the correspondence between quantum $SL(2,C)$ and Lorentz groups developped in [3].\\ The unimodularity of $M_{\alpha}^{~\beta}$ is expressed by $\varepsilon_{\alpha\beta}M_{\gamma}^{~\alpha} M_{\delta}^{~\beta}= \varepsilon_{\gamma\delta}I_{\cal A}$, $\varepsilon^{\gamma\delta}M_{\gamma}^{~\alpha} M_{\delta}^{~\beta}= \varepsilon^{\alpha\beta}I_{\cal A}$, and $\varepsilon_{\dot{\alpha}\dot{\beta}}M_{\dot{\gamma}}^{~\dot{\alpha}} M_{\dot{\delta}}^{~\dot{\beta}}= \varepsilon_{\dot{\gamma}\dot{\delta}}I_{\cal A}$, $\varepsilon^{\dot{\gamma}\dot{\delta}}M_{\dot{\gamma}}^{~\dot{\alpha}} M_{\dot{\delta}}^{~\dot{\beta}}= \varepsilon^{\dot{\alpha}\dot{\beta}}I_{\cal A}$ where $I_{\cal A}$ is the unity of the $\star$ algebra ${\cal A}$ generated by $M_{\alpha}^{~\beta}$ and the spinor metric $\varepsilon_{\alpha\beta}$ and its inverse $\varepsilon^{\alpha\beta}~(\varepsilon_{\alpha\delta} \varepsilon^{\delta\beta}=\delta_{\alpha}^{\beta}= \varepsilon^{\beta\delta}\varepsilon_{\delta\alpha})$ satisfy $(\varepsilon_{\alpha\beta})^{\star} =\varepsilon_{\dot{\beta}\dot{\alpha}}$ and $(\varepsilon^{\alpha\beta})^{\star}=\varepsilon^{\dot{\beta}\dot{\alpha}}$. The commutation rules are given by $M_{\alpha}^{~\gamma}(f_{\pm\gamma}^{~~\beta}\star a) = (a \star f_{\pm\alpha}^{~~\gamma})M_{\gamma}^{~\beta}$ and $M_{\dot{\alpha}}^{~\dot{\gamma}}(f_{\pm\dot{\gamma}}^{~~\dot{\beta}}\star a)= (a \star f_{\pm\dot{\alpha}}^{~~\dot{\gamma}})M_{\dot{\gamma}}^{~\dot{\beta}}$ , for any $a\in \cal A$, where $f_{\pm\alpha}^{~~\beta}$ and $f_{\pm\dot{\alpha}}^{~~\dot{\beta}}$ $\in \cal A'$, dual to the Hopf algebra $\cal A$, are linear functionals $:\cal A \rightarrow \cal C$. The convolution product is defined as $f\star a=(id \otimes f)\Delta(a)$ and $a\star f=(f\otimes id)\Delta(a)$. The subalgebra ${\cal A}'_{0} \subset {\cal A}'$ generated by $f_{\pm\alpha}^{~~\beta}$ and $f_{\pm\dot{\alpha}}^{~~\dot{\beta}}$ is a Hopf algebra acting on the genrators of $\cal A$ as $f_{\pm\alpha}^{~~\beta}(M_{\gamma}^{~\delta})= a^{\mp \frac{1}{2}}R^{\pm\delta\beta}_{~\alpha\gamma}$ and $f_{\pm\dot{\alpha}}^{~~\dot{\beta}}(M_{\dot{\gamma}}^{~\dot{\delta}})= a^{\mp \frac{1}{2}} R^{\pm\dot{\delta}\dot{\beta}}_{~\dot{\alpha}\dot{\gamma}}$ [4] where $a\not=0$ is a real number satisfying $\varepsilon_{\alpha\beta}\varepsilon^{\alpha\beta} = -(a + a^{-1})=-Q$. The R-matrices $R^{\pm\delta\beta}_{~\alpha\gamma}= \delta^{\delta}_{\alpha}\delta^{\beta}_{\gamma} + a^{\pm 1} \varepsilon^{\delta\beta}\varepsilon_{\alpha\gamma}~( R^{\pm\dot{\delta}\dot{\beta}}_{~\dot{\alpha}\dot{\gamma}}= \delta^{\dot{\delta}}_{\dot{\alpha}}\delta^{\dot{\beta}}_{\dot{\gamma}} + a^{\pm 1}\varepsilon^{\dot{\delta}\dot{\beta}} \varepsilon_{\dot{\alpha}\dot{\gamma}})$ satisfy $R^{\pm\delta\beta}_{~\alpha\gamma}R^{\mp\rho\sigma}_{~\delta\beta}= \delta^{\rho}_{\alpha}\delta^{\sigma}_{\gamma}~( R^{\pm\dot{\delta}\dot{\beta}}_{~\dot{\alpha}\dot{\gamma}} R^{\mp\dot{\rho}\dot{\sigma}}_{~\dot{\delta}\dot{\beta}}= \delta^{\dot{\rho}}_{\dot{\alpha}}\delta^{\dot{\sigma}}_{\dot{\gamma}})$, the Hecke conditions $(R^{\pm}+a^{\pm 2})(R^{\pm}-1)=0$ and the Yang-Baxter equations [5].\\ Besides commutation rules between undotted or dotted generators only, we need to specify commutation rules between undotted and dotted generators which are controlled by $R$-matrices $f_{\pm\alpha}^{~\beta}(M_{\dot{\gamma}}^{~\dot{\delta}})= R^{\pm\dot{\delta}\beta}_{~\alpha\dot{\gamma}}$ which we assume they take the form of the R matrix of the quantum $SU(2)$ group. More precisely, the generators $M_{\alpha}^{~\beta}$ must satisfy the unitarity condition, $M_{\dot{\alpha}}^{~\dot{\beta}}=S(M_{\beta}^{~\alpha})$ and $M_{\alpha}^{~\beta}=S^{-1}(M_{\dot{\beta}}^{~\dot{\alpha}})$, when they are considered as belonging to a range of functionals $f_{\pm\alpha}^{~~\beta}$, $f_{\pm\alpha}^{~~\beta}(M_{\dot{\gamma}}^{~\dot{\delta}})= R^{\pm\dot{\delta}\beta}_{~\alpha\dot{\gamma}}= f_{\pm\alpha}^{~~\beta}(S(M_{\delta}^{~\gamma}))=a^{\pm\frac{1}{2}} R^{\mp\beta\gamma}_{~\delta\alpha}$ and $f_{\pm\alpha}^{~~\beta}(S(M_{\dot{\gamma}}^{~\dot{\delta}}))= R^{\mp\beta\dot{\delta}}_{~\dot{\gamma}\alpha}= \varepsilon_{\dot{\gamma}\dot{\rho}}f_{\pm\alpha}^{~~\beta}(M_{\dot{\sigma}} ^{~\dot{\rho}})\varepsilon^{\dot{\sigma}\dot{\delta}}$. In this case, the spinor metrics must satisfy an additional condition, $\varepsilon_{\alpha\beta}=-\varepsilon^{\beta\alpha}$ and $\varepsilon_{\dot{\alpha}\dot{\beta}}=- \varepsilon^{\dot{\beta}\dot{\alpha}}$, required by consistency condition between the unitarity and the unimodularity of the quantum $SU(2)$ group generators. The functionals $f_{\pm\alpha}^{~~\beta}$ and $f_{\pm\dot{\alpha}}^{~~\dot{\beta}}$ satisfy the conditions $f_{\pm\dot{\alpha}}^{~~\dot{\beta}} =\tilde{f}_{\pm\beta}^{~~\alpha}$ and $\tilde{f}_{\pm\dot{\alpha}}^{~~\dot{\beta}} = f_{\pm\beta}^{~~\alpha}$ where $f_{\pm\beta}^{~~\alpha}=\tilde{f}_{\pm\beta}^{~~\alpha}\circ S = \varepsilon^{\alpha\delta}\tilde{f}_{\pm\delta}^{~~\gamma} \varepsilon_{\gamma\beta}$ and $f_{\pm\dot{\alpha}}^{~~\dot{\beta}}= \tilde{f}_{\pm\dot{\alpha}}^{~~\dot{\beta}}\circ S^{-1}= \varepsilon_{\dot{\alpha}\dot{\gamma}} \tilde{f}_{\pm\dot{\delta}}^{~~\dot{\gamma}} \varepsilon^{\dot{\delta}\dot{\beta}}$. These functionals also satisfy $(f_{\pm\beta}^{~~\alpha}(a))^{\star} = f_{\mp\dot{\alpha}}^{~~\dot{\beta}}(S(a^{\star}))$ and $(f_{\pm\dot{\beta}}^{~~\dot{\alpha}}(a))^{\star} = f_{\mp\alpha}^{~~\beta}(S(a^{\star}))$.\\ With these hypothesis on the commutation rules, it is shown in [3] that the generators $\Lambda_{N}^{~M}$ of quantum Lorentz group may be written in terms of those of quantum $SL(2,C)$ group as $\Lambda_{N}^{~M} =\frac{1}{Q}\varepsilon_{\dot{\gamma}\dot{\delta}} \overline{\sigma}_{N}^{~\dot{\delta}\alpha}M_{\alpha}^{~\sigma} \sigma^{M}_{~\sigma\dot{\rho}}M_{\dot{\beta}}^{~\dot{\rho}} \varepsilon^{\dot{\gamma}\dot{\beta}}$. They are real, $(\Lambda_{N}^{~M})^{\star}=\Lambda_{N}^{~M}$, and generate a Hopf algebra $\cal L$ endowed with a coaction $\Delta$, a counit $\varepsilon$ and an antipode $S$ acting as $\Delta(\Lambda_{N}^{~M}) = \Lambda_{N}^{~K}\otimes \Lambda_{K}^{~M}$, $\varepsilon(\Lambda_{N}^{~M})= \delta_{N}^{M}$ and $S(\Lambda_{N}^{~M})= G_{\pm NK}\Lambda_{L}^{~K}G_{\pm}^{LM}$ respectively. $G_{\pm}^{NM}$ is an invertible and hermitian quantum metric. It may be expressed in terms of the four matrices $\sigma^{N}_{\alpha\dot{\beta}}~(N=0,1,2,3)$ as: \begin{eqnarray*} G_{\pm}^{~IJ} =\frac{1}{Q}Tr(\sigma^{I} \overline{\sigma}_{\pm}^{J}) = \frac{1}{Q}\varepsilon^{\alpha \nu}\sigma^{I}_{~\alpha\dot{\beta}} \overline{\sigma}_{\pm}^{J\dot{\beta}\gamma}\varepsilon_{\gamma\nu}= \frac{1}{Q}Tr(\overline{\sigma}_{\pm}^{I} \sigma^{J})= \frac{1}{Q} \varepsilon_{\dot{\nu}\dot{\gamma}} \overline{\sigma}_{\pm}^{I\dot{\gamma}\alpha} \sigma^{J}_{~\alpha\dot{\beta}} \varepsilon^{\dot{\nu}\dot{\beta}} \end{eqnarray*} where $\sigma^{n}_{\alpha\dot{\beta}}~(n=1,2,3)$ are the usual Pauli matrices, $\sigma^{0}_{\alpha\dot{\beta}}$ is the identity matrix and $\overline{\sigma}_{\pm}^{I\dot{\alpha} \beta} = \varepsilon^{\dot{\alpha} \dot{\lambda}}R^{\mp \sigma \dot{\rho}}_{~~\dot{\lambda}\nu} \varepsilon^{\nu \beta}\sigma^{I}_{~\sigma \dot{\rho}}$. The inverse of the Minkowskian metric may be written under the form $G_{\pm IJ}=\frac{1}{Q}Tr(\overline{\sigma}_{J}\sigma_{\pm I})= \frac{1}{Q} \varepsilon_{\dot{\nu}\dot{\gamma}} \overline{\sigma}_{J}^{~\dot{\gamma}\alpha} \sigma_{\pm I\alpha\dot{\beta}} \varepsilon^{\dot{\nu}\dot{\beta}}$ where $\sigma_{\pm I\alpha \dot{\beta}} = G_{\pm IJ}\sigma^{J}_{\alpha\dot{\beta}}$. The form of the antipode of $\Lambda_{N}^{~M}$ guarantees the orthogonality condition on the generators of quantum Lorentz group as: \begin{eqnarray} G_{\pm NM}\Lambda_{L}^{~N}\Lambda_{K}^{~M} = G_{\pm LK}I_{\cal L}~~and~~ G_{\pm}^{~LK}\Lambda_{L}^{~N}\Lambda_{K}^{~M} = G_{\pm}^{~NM}I_{\cal L} \end{eqnarray} where $I_{{\cal L}}$ is the unity of the algebra $\cal L$.\\ The completeness relations are given by $\sigma^{I~\dot{\beta}}_{\alpha} \overline{\sigma}_{I\dot{\rho}}^{~~~\sigma}= Q\delta^{\sigma}_{\alpha}\delta_{\dot{\rho}}^{\dot{\beta}}$ and $\sigma_{\pm I~\dot{\beta}}^{~~\alpha} \overline{\sigma}^{I\dot{\rho}}_{\pm~\sigma}= Q\delta_{\sigma}^{\alpha}\delta^{\dot{\rho}}_{\dot{\beta}}$ where the undotted and dotted spinorial indices are raised and lowered as $\sigma^{I\alpha}_{~~\dot{\beta}} = \sigma^{I}_{~\rho\dot{\beta}} \varepsilon^{\rho\alpha}$ and $\sigma^{I~\dot{\beta}}_{~\alpha}= \varepsilon^{\dot{\beta}\dot{\rho}}\sigma^{I}_{\alpha\dot{\rho}}$. These completeness relations may be used to convert a vector to a bispinor and vice versa \begin{eqnarray*} X_{\alpha\dot{\beta}} = X_{I}\sigma^{I}_{\alpha\dot{\beta}}\Leftrightarrow X_{I}= \frac{1}{Q}\varepsilon^{\alpha\nu}X_{\alpha\dot{\beta}} \overline{\sigma}_{\pm}^{J\dot{\beta}\delta}\varepsilon_{\delta\nu} G_{\pm JI}~~or ~~ X_{I}= \frac{1}{Q}\varepsilon_{\dot{\nu}\dot{\beta}} \overline{\sigma}_{I}^{~\dot{\beta}\alpha}X_{\alpha\dot{\delta}} \varepsilon^{\dot{\nu}\dot{\delta}} \end{eqnarray*} where $X_{I}$ are real elements of the right invariant basis of a bimodule $\cal M$ over the algebra $\cal L$ which transform under the left coaction $\Delta_{L}:{\cal M} \rightarrow {\cal L}\otimes {\cal M}$ as \begin{eqnarray} \Delta_{L}(X_{I})=\Lambda_{I}^{~K}\otimes X_{K}. \end{eqnarray} The functional $f_{\alpha}^{~\beta}$ of $SL(2,C)$ induce that of Lorentz group $F_{\pm L}^{~~K}: {\cal L} \rightarrow C$ given by $F_{\pm L}^{~~K} = \frac{1}{Q}(\tilde{f}_{\mp \dot{\beta}}^{~~~\dot{\alpha}} \overline{\sigma}_{L\dot{\alpha}}^{~~~\delta} \star f_{\pm\delta}^{~~~\gamma} \sigma^{K~\dot{\beta}}_{~\gamma})$. They controle the noncommutatitivity between elements of $\cal L$ and $X_{I}$ as \begin{eqnarray} \Lambda_{L}^{~I}(F_{\pm I}^{~~K} \star a) &=& (a \star F_{\pm L}^{~~I})\Lambda_{I}^{~K} ~,\\ X_{\pm L}a=(a \star F_{\pm L}^{~~K})X_{\pm K}~&and&~X_{(a)L}X_{(b)K}= F_{(b)K}^{~~~~N}(S(\Lambda_{L}^{~M}))X_{(b)N}X_{(a)M} \end{eqnarray} where $a\in \cal L$ and the indices $a,b=\pm$. These functionals satisfy $F_{\pm L}^{~~K}(\delta_{N}^{M})=\delta_{L}^{K}\delta_{N}^{M}$, $(F_{\pm L}^{~~K}(a))^{\star}=F_{\pm L}^{~~K}(S(a^{\star}))$ for any $a\in \cal L$ and the relations $G_{\pm}^{~MN}F_{\pm N}^{~~L} \star F_{\pm M}^{~~K}(a) = G_{\pm}^{~KL} \varepsilon(a)$ and $G_{\pm KL}F_{\pm N}^{~~L} \star F_{\pm M}^{~~K}(a) = G_{\pm MN}\varepsilon(a)$ which imply that the length of the quantum four-vector, $G^{LK}X_{L}X_{K}$, is bi-invariant, central and real. Since it commutes with everything, it is of the form $G^{LK}X_{L}X_{K}=-\tau^{2} I_{\cal L}$ where $\tau^{2}$ is a real number.\\ We can also show that the quantum symmetrization of the Minkowski metric is given by ${\cal R}^{\pm NM}_{~KL}G_{\pm}^{KL} = G_{\pm}^{NM}$ and ${\cal R}^{\pm NM}_{~KL}G_{\pm NM} = G_{\pm KL}$ where ${\cal R}^{\pm NM}_{~KL} = F_{\pm K}^{~~M}(\Lambda_{L}^{~N})$ satisfy the Yang-Baxter equations and the cubic Hecke conditions $({\cal R}^{\pm} + a^{\pm2})({\cal R}^{\pm} + a^{\mp2}) ({\cal R}^{\pm} - 1)=0$.\\ In following we shall consider the right invariant basis $X_{I}=X_{+I}$ of the bimodule algebra ${\cal M}$as a quantum coordinate system of the Minkowski space-time $\cal M$ provided with the metric $G^{IJ}=G_{+}^{~IJ}$. $X_{0}$ represents the time operator and $X_{i}~(i=1,2,3)$ represent the space coordinate operators.\\ To make the explicit calculation of the different commutation rules, we take an adequate choice of Pauli hermitian matrices of the form \begin{eqnarray*} \sigma^{0}_{\alpha \dot{\beta}} =\left(\begin{array}{cc} 1 & 0\\ 0 & 1 \end{array} \right)~~,~~ \sigma^{1}_{\alpha \dot{\beta}} =\left(\begin{array}{cc} 0 & 1\\1 & 0 \end{array} \right)~~,~~ \sigma^{2}_{\alpha \dot{\beta}} =\left(\begin{array}{cc} 0 & -i\\ i & 0 \end{array} \right)~~,~~ \sigma^{3}_{\alpha \dot{\beta}} =\left(\begin{array}{cc} q & 0\\ 0 &-q^{-1} \end{array} \right) \end{eqnarray*} and the spinorial metrics $\varepsilon_{\alpha\beta}=-\varepsilon^{\alpha\beta}= \varepsilon_{\dot{\beta}\dot{\alpha}}=-\varepsilon^{\dot{\beta}\dot{\alpha}} =\left(\begin{array}{cc} 0& -q^{-\frac{1}{2}}\\q^{\frac{1}{2}}&0 \end{array} \right)$ which presents an advantage arising from the fact that, in this representation, we have $\overline{\sigma}_{0}^{~\dot{\alpha}\beta} = -\sigma^{0}_{\alpha \dot{\beta}}=-\delta_{\alpha}^{\beta}$, $\overline{\sigma}_{N\dot{\alpha}\alpha}=\overline{\sigma}_{N\dot{1}1} + \overline{\sigma}_{N\dot{2}2}=-(q+q^{-1})\delta_{N}^{0}=-Q\delta_{N}^{0} $ and $\sigma^{N\alpha\dot{\alpha}}=Q\delta_{0}^{N}$. We shall see in the next section that these properties lead directly to the restriction of the quantum lorentz group to the quantum subgroup of the three dimensional space rotations by restricting the quantum $SL(2,C)$ group to the $SU(2)$ group. These representations also give a form of the quantum metric $G^{NM}$ into two independent blocks, one for the time component $X_{0}$ and the other for space components $X_{k}~(k=1,2,3)$. The non vanishing elements of the metric are $G^{00}=-q^{-\frac{3}{2}}$, $G^{11}=G^{22}=G^{33}=q^{\frac{1}{2}}$, $G^{12}=-G^{21}=-iq^{\frac{1}{2}}\frac{q-q^{-1}}{Q}$ and the non vanishing elements of its inverse are $G_{00}=-q^{\frac{3}{2}}$, $G_{11}=G_{22}=q^{-\frac{1}{2}}\frac{Q^{2}}{4}$, $G_{33}=q^{-\frac{1}{2}}$, $G_{12}=-G_{21}=iq^{-\frac{1}{2}}\frac{(q-q^{-1})Q}{4}$. In the classical limit $q=1$, this metric reduces to the classical Minkowski metric with signature $(-,+,+,+)$.\\ \section{\bf The lifetime dilatation in the quantum Minkowski space-time} Before to show how to derive $q$-analog physical properties of the usual special relativity, let us investigate the properties of the $\cal R$ matrix and its consequences on the different commutation rules. Let us recall that as a range of the functionals $f_{\alpha}^{~\beta}$ the generators $M_{\alpha}^{~\beta}$ of the Hopf algebra $\cal A$ satisfy unitarity condition. Hence as a range of the functionals $F_{N}^{~M}$, the generators of $\cal L$ have the following properties \begin{eqnarray} \Lambda_{N}^{~0} =\frac{1}{Q} \overline{\sigma}_{N\dot{\gamma}}^{~~~\alpha}M_{\alpha}^{~\sigma} \sigma^{0}_{~\sigma\dot{\rho}}S(M_{\rho}^{~\beta}) \varepsilon^{\dot{\gamma}\dot{\beta}} &=&\frac{1}{Q} \overline{\sigma}_{N\dot{\gamma}}^{~~~\alpha} \varepsilon^{\dot{\gamma}\dot{\alpha}}\nonumber\\ =-\frac{1}{Q}\overline{\sigma}_{N\dot{\gamma}\gamma}&=& \delta_{N}^{0} \end{eqnarray} and \begin{eqnarray} \Lambda_{0}^{~M} =\frac{1}{Q}\varepsilon_{\dot{\gamma}\dot{\delta}} \overline{\sigma}_{0}^{~\dot{\delta}\alpha}M_{\alpha}^{~\sigma} \sigma^{M}_{~\sigma\dot{\rho}}S(M_{\rho}^{~\beta}) \varepsilon^{\dot{\gamma}\dot{\beta}} &=& -\frac{1}{Q}\varepsilon^{\alpha\gamma} M_{\alpha}^{~\sigma}\sigma^{M}_{~\sigma\dot{\rho}}\varepsilon_{\rho\delta} M_{\gamma}^{~\delta}=\nonumber\\ \frac{1}{Q} \varepsilon^{\dot{\delta}\dot{\rho}} \sigma^{M}_{~\sigma\dot{\rho}}\varepsilon^{\sigma\delta} &=& \frac{1}{Q} \sigma^{M\delta\dot{\delta}}=\delta_{0}^{M} \end{eqnarray} from which, we get \begin{eqnarray} F_{N}^{~M}(\Lambda_{P}^{~0}) =\delta_{N}^{M}\delta_{P}^{0} = {\cal R}^{0M}_{NP},~and~F_{N}^{~M}(\Lambda_{0}^{~Q}) = \delta_{N}^{M}\delta_{0}^{Q}={\cal R}^{QM}_{N0}. \end{eqnarray} The quantum symmetrization of the coordinates may be written, from the right relation of (4), as \begin{eqnarray} X_{L}X_{K} = {\cal R}^{NM}_{LK}X_{N}X_{M} \end{eqnarray} leading for $K=0$ to \begin{eqnarray} X_{L}X_{0} - {\cal R}^{NM}_{L0}X_{N}X_{M}=X_{0}X_{L} \end{eqnarray} which shows that the time coordinate operator commutes with the space coordinate operator. After a straighforward computation (8) gives the commutation relations between space coordinates as: \begin{eqnarray} X_{3}Z - q^{2}ZX_{3}=(q - q^{-1})X_{0}Z,\\ X_{3}\overline{Z} - q^{-2}\overline{Z}X_{3}= - q^{-2}(q - q^{-1})X_{0}\overline{Z},\\ Z\overline{Z}-\overline{Z}Z = (q^{2} - q^{-2})X_{3}^{2} + q^{-1}(q^{2} - q^{-2})X_{0}X_{3} \end{eqnarray} where $Z=X_{1}+iX_{2}$ and $\overline{Z}=X_{1}-iX_{2}$. The four-vector length $G^{IJ}X_{I}X_{J} = -\tau^{2}I$ may be written of the form \begin{eqnarray} q^{-\frac{3}{2}}X_{0}^{2} - \frac{q^{\frac{3}{2}}}{Q}Z\overline{Z} - \frac{q^{-\frac{1}{2}}}{Q}\overline{Z}Z - q^{\frac{1}{2}}X_{3}^{2} = \tau^{2}. \end{eqnarray} Note that by redefining the Minkowski space-time coordinates as $C=qX_{0}-X_{3}$, $D=q^{-1}X_{0}+X_{3}$, $A=Z$ and $B=\overline{Z}$ we recover the commutation relations and the four-vector length given in Ref.[2,6,7].\\ As a consequence of (7), the relation (3) gives, for $a=\Lambda_{N}^{~0}$, \begin{eqnarray} \Lambda_{L}^{~I}\Lambda_{N}^{~P}F_{I}^{~K}(\Lambda_{P}^{~0})= \Lambda_{L}^{~K}\Lambda_{N}^{~0}=F_{L}^{~I}(\Lambda_{N}^{~P}) \Lambda_{P}^{~0}\Lambda_{I}^{~K}={\cal R}^{PI}_{LN}\Lambda_{P}^{~0} \Lambda_{I}^{~K} \end{eqnarray} which reduces to \begin{eqnarray} \Lambda_{L}^{~0}\Lambda_{N}^{~0}={\cal R}^{PI}_{LN}\Lambda_{P}^{~0} \Lambda_{I}^{~0} \end{eqnarray} for $K=0$ and to \begin{eqnarray} \Lambda_{L}^{~K}\Lambda_{0}^{~0}=\Lambda_{0}^{~0}\Lambda_{L}^{~K} \end{eqnarray} for $N=0$. The commutation relations between the Minkowski space-time coordinates $X_{L}$ and the generators of the quantum Lorentz group $\Lambda_{N}^{~M}$ are given by the left relation of (4) which gives for $a=\Lambda_{N}^{~0}$ \begin{eqnarray} X_{L}\Lambda_{N}^{~0} = F_{L}^{~K}(\Lambda_{N}^{~P})\Lambda_{P}^{~0} X_{K} = {\cal R}^{PK}_{LN}\Lambda_{P}^{~0}X_{K}. \end{eqnarray} This relation reduces to \begin{eqnarray} X_{L}\Lambda_{0}^{~0} = F_{L}^{~K}(\Lambda_{0}^{~P})\Lambda_{P}^{~0} X_{K} = \Lambda_{0}^{~0}X_{L} \end{eqnarray} for $N=0$.\\ We are now ready to show that the above commutation relations suffice to construct the physical states and the different $q$-deformed observables derived from the quantum boost of particles at rest. Since the coordinates and their transformations are operators, we consider, for the usual quantum mechanics, that the evolution of a free particle in the noncommutative special relativity is described by Hilbert space states which are common eingenstates of a set of commuting elements of the quantum algebras ${\cal M}$ and ${\cal L}$. The corresponding eingenvalues give the measurable quantities tied to the evolution of this particle.\\ First we see from the relations (16) and (18) that $\Lambda_{0}^{~0}$ commutes with the algebras $\cal L$ and $\cal M$, then it is a real c-number. Therefore, as in textbooks of special relativity we set $\Lambda_{0}^{~0}=\gamma I_{\cal L} $ and $\Lambda_{I}^{~0}=\gamma V_{I}$ where $V_{0}=I_{\cal L}$ and $V_{i}$ are the components of the velocity operator. Due to the fact that $\gamma$ is a real c-number and $\Lambda_{I}^{~0}$ are real operators, the components $V_{i}$ of the velocity are real operators. From (15) we see that $\Lambda_{I}^{~0}$ fulfil the same commutation relations (8) of the coordinates $X_{I}$, then $V_{I}$ satisfy the commutations rules $V_{L}V_{K}=R_{LK}^{NM}V_{N}V_{M}$ which give the same relations (10-12) by replacing $X_{i}$ by $V_{i}$ and $X_{0}$ by $1$. With these notations, the orthogonality relations (1) give \begin{eqnarray} G^{IJ}\Lambda_{I}^{~0}\Lambda_{J}^{~0} = \gamma^{2}(G^{00} + G^{ij}V_{i}V_{j})= G^{00} \end{eqnarray} leading to \begin{eqnarray} \gamma =(1 - |v|_{q}^{2})^{-\frac{1}{2}}. \end{eqnarray} Since $\gamma$ is a real c-number, the length of the velocity, $-\frac{G^{ij}}{G^{00}}V_{i}V_{j} = |v|_{q}^{2}$, is also a real c-number. Then both $\gamma$ and $|V|_{q}$ can be measured exactly. $\Lambda_{0}^{~0}=(\varepsilon^{\alpha\delta}M_{\alpha}^{~\sigma}) (\varepsilon^{\alpha\delta}M_{\alpha}^{~\sigma})^{\star}$ is a real and positive operator, then $\gamma >0$. From (20) we see that the upper limit of the velocity is $1$ (the light velocity). This point will be discussed in more details in the next section.\\ Now, we show how the states transform under a Lorentz group. To do that, we observe that we are in presence of two quantum algebras, $\cal L$ corresponds to the quantum symmetries and $\cal M$ is generated by the quantum coordinates of the Minkowski space-time. Since the coordinates transform through the left coaction (2) as a tensor product ${\cal L} \otimes {\cal M}$, we consider the space states of positive norm as a tensor product ${\cal H}_{\cal L} \otimes {\cal H}_{\cal M}$ of two Hilbert spaces in which the algebras $\cal L$ and $\cal M$ act respectively (space of representations of the coordinate transformations). Then in this formalism we have to construct bases for ${\cal H}_{\cal M}$ and ${\cal H}_{\cal L}$ which are eingenstate of the set of commuting elements of algebras $\cal L$ and $\cal M$ separately.\\ From the commutation rules of the coordinates (9-12), we see that the set of commuting operators are the proper time $\tau^{2}$, the time component $X_{0}$ and one of the space components, for example $X_{3}$. Then in a fixed reference system, a particle is described by a state belonging to the Hilbert space ${\cal H}_{\cal M}$ labeled by the eigenvalues $\tau^{2}$, $t$ and $x_{3}$, $|{\cal P}\rangle=|t,x_{3},\tau^{2}\rangle$ eingenstate of $X_{0}$, $X_{3}$ and $\tau^{2}$ respectively \begin{eqnarray} X_{0}|{\cal P}\rangle=t |{\cal P}\rangle~~,~~ X_{3}|{\cal P}\rangle= x_{3} |{\cal P}\rangle~~,~~ \tau^{2}|{\cal P}\rangle=\tau^{2}|{\cal P}\rangle. \end{eqnarray} This state describes the evolution of a particle seen by an observer $O$ in a coordinate system $X_{N}$. For a second observer $O'$ connected with the observer $O$ by a quantum Lorentz transformation, the coordinate system tied to $O'$ is given by $X'_{N}=\Delta_{L}(X_{N})=\Lambda_{N}^{~M}\otimes X_{M}$ which fulfil the same commutation rules (9-12) as $X_{N}$. Then the state $|{\cal P}'\rangle$, describing the same particle seen by the observer $O'$ is also labeled by the eigenvalues of $\tau^{2}$, $X'_{0}$ and $X'_{3}$ as $|{\cal P}'\rangle=|t',x'_{3},\tau^{2}\rangle \in {\cal H}_{\cal M}$ satisfying: \begin{eqnarray} X'_{0}|{\cal P}'\rangle=t' |{\cal P}'\rangle~~,~~ X'_{3}|{\cal P}'\rangle= x'_{3} |{\cal P}'\rangle~~,~~ \tau^{2}|{\cal P}'\rangle=\tau^{2}|{\cal P}'\rangle. \end{eqnarray} Since the coordinates transform with tensorial products as $X'_{N}=\Lambda_{N}^{~M}\otimes X_{M}$ it is natural to suppose that the Hilbert state $|{\cal P}\rangle$ transforms into $|{\cal P}'\rangle$ as \begin{eqnarray} |{\cal P}'\rangle=|sym_{q}\rangle \otimes |{\cal P}\rangle \end{eqnarray} where $|sym_{q}\rangle$ is a basis of the Hilbert space ${\cal H}_{\cal L}$ where the generators $\Lambda_{N}^{~M}$ of the quantum Lorentz group are represented. Note that:\\ - The state transformation (23) are different from those of the usual quantum mechanics where the states transform with the unitary transformations $U(G)$ of the classical group $G$ as $|\Psi '\rangle=U(G)|\Psi\rangle$.\\ - At this stage there is no relation between the eigenvalues of space coordinates and those of Lorentz group as velocities. So (23) may be regarded as the quantum analog of a point in the configuration space of the classical mechanics.\\ Under the quantum Lorentz transformation, the coordinates $X'_{N}$ act on $|{\cal P}'\rangle$ as: \begin{eqnarray} X'_{0}|{\cal P}'\rangle = (\Lambda_{0}^{~0}\otimes X_{0})|{\cal P}'\rangle + (\Lambda_{0}^{~k}\otimes X_{k})|{\cal P}'\rangle= \Lambda_{0}^{~0}|sym_{q}\rangle\otimes X_{0}|{\cal P}\rangle + \Lambda_{0}^{~k}|sym_{q}\rangle\otimes X_{k}|{\cal P}\rangle,\\ X'_{i}|{\cal P}'\rangle = (\Lambda_{i}^{~0}\otimes X_{0})|{\cal P}'\rangle + (\Lambda_{i}^{~k}\otimes X_{k})|{\cal P}'\rangle= \Lambda_{i}^{~0}|sym_{q}\rangle\otimes X_{0}|{\cal P}\rangle + \Lambda_{i}^{~k}|sym_{q}\rangle\otimes X_{k}|{\cal P}\rangle. \end{eqnarray} In the following we consider the simplest case where a particle at rest is boosted with a velocity $v$. Let $|{\cal P}_{0}\rangle =|t,0,\tau^{2}\rangle$ a state describing a particle at rest. This state satisfy beside $X_{0}|t,0,\tau^{2}\rangle=t|t,0,\tau^{2}\rangle$, $\tau^{2}|t,0,\tau^{2}\rangle=\tau^{2}|t,0,\tau^{2}\rangle$ and $X_{3}|t,0,\tau^{2}\rangle=0|t,0,\tau^{2}\rangle$ two additionnal conditions \begin{eqnarray} X_{1}|{\cal P}\rangle= 0 |{\cal P}\rangle~~,~~ X_{2}|{\cal P}\rangle= 0 |{\cal P}\rangle. \end{eqnarray} The latter relations are possible because of the homogeneous feature of the commutation rules (10-12) and do not desagree with (13). In the next section, we shall show that this rest state is unique. By using the properties (26) of the rest state, the transformations (24-25) reduce to \begin{eqnarray} X'_{0}|{\cal P}'\rangle = t'|{\cal P}'\rangle = \Lambda_{0}^{~0}|sym_{q}\rangle\otimes X_{0}|{\cal P}_{0}\rangle= \gamma|sym_{q}\rangle\otimes t|{\cal P}_{0}\rangle= \gamma t'|{\cal P}'\rangle,\\ X'_{i}|{\cal P}'\rangle = \Lambda_{i}^{~0}|sym_{q}\rangle\otimes X_{0}|{\cal P}_{0}\rangle = V_{i}\gamma|sym_{q}\rangle\otimes t|{\cal P}_{0}\rangle= V_{i}\gamma t|{\cal P}'\rangle= V_{i}t'|{\cal P}'\rangle. \end{eqnarray} These transformations require the knowledg of the commutation rules between $\Lambda_{N}^{~0}=V_{N}\gamma$ only, which can be deduced from (9-12). Then as the coordinate system, the state describing the quantum boost are labeled by the length $|v|_{q}^{2}$ or $\gamma$ and its component $v_{3}$. And, therefore the transformed Hilbert space state may be written under the form of tensor product of the rest state $|{\cal P}_{0}\rangle$ and the boost state labelled by $v_{3}$ and the length of the velocity $|v|_{q}^{2}$ eigenvalues of the common eigenstate $|v_{3},|v|_{q}^{2}\rangle$ of $V_{3}$ and $-\frac{G^{ij}}{G^{00}}V_{i}V_{j}$ as: \begin{eqnarray} |{\cal P}'\rangle =|v_{3},|v|_{q}^{2}\rangle \otimes |{\cal P}_{0}\rangle. \end{eqnarray} On this state, the transformation (27-28) lead to \begin{eqnarray} X'_{0}|{\cal P}'\rangle = t'|{\cal P}'\rangle = (\gamma\otimes t)|{\cal P}'\rangle~~and~~ X'_{3}|{\cal P}'\rangle = x'_{3}|{\cal P}'\rangle = (v_{3}\otimes\gamma t)|{\cal P}'\rangle= v_{3}t'|{\cal P}'\rangle \end{eqnarray} which show the usual relation $x'_{3}=v_{3}t$ between the coordinate $x_{3}$ and the component $v_{3}$ of the velocity and the time for a free particle moving with the velocity $|v|_{q}$. Note that we have assumed $t\geq 0$, since $\gamma\geq 0$ then $t'=\gamma t\geq 0$. On the other hand the combination of the invariance of the proper time $\tau^{2}$ with (26-28) and their conjugate ($\langle{\cal P}|X'_{i}=\langle{\cal P}|V_{i}t$) gives \begin{eqnarray*} \langle{\cal P}'|G^{IJ}X'_{I}X'_{J}|{\cal P}'\rangle = \langle{\cal P}'|G^{00}(X'^{2}_{0} &+& \frac{G^{ij}}{G^{00}}X'_{i}X'_{j}) |{\cal P}'\rangle = \langle{\cal P}'|G^{00}(t'^{2}+\frac{G^{ij}}{G^{00}}V_{i}V_{j}t'^{2}) |{\cal P}'\rangle =\\ \langle{\cal P}'|G^{00}(t'^{2}-|v|_{q}^{2}t'^{2})|{\cal P}'\rangle&=& G^{00}(t'^{2}-|v|_{q}^{2}t'^{2})=\\ \langle v_{3},|v|_{q}^{2}|v_{3},|v|_{q}^{2}\rangle\otimes \langle{\cal P}_{0}|G^{IJ}X_{I}X_{J}|{\cal P}_{0}\rangle &=& \langle v_{3},|v|_{q}^{2}|v_{3},|v|_{q}^{2}\rangle\otimes \langle{\cal P}_{0}|G^{00}t^{2}|{\cal P}_{0}\rangle = \\ I&\otimes& G^{00}t^{2} \end{eqnarray*} leading to the $q$-deformed lifetime dilatation in the noncommutative special relativity \begin{eqnarray} t'=\frac{t}{(1-|v|_{q}^{2})^{\frac{1}{2}}}. \end{eqnarray} In the semi classical limit $q\simeq 1\pm \kappa + O(\kappa^{2})$, where $|v|_{q=0}=|v_{cl}|$ denotes the classical velocity, the correction to the classical formula of the lifetime dilatation of unstable particles $\Delta(t_{cl})$ is given by \begin{eqnarray} \Delta(t'_{q}) \simeq \Delta(t_{cl}) \mp \kappa\Delta(t_{cl})\frac{|v_{cl}|^{2}} {1-|v_{cl}|^{2}} + O(\kappa^{2}). \end{eqnarray} Note that we can also consider the bicovariant bimodule ${\cal M}$ over ${\cal L}$ where, instead of space-time coordinates $X_{N}$, we take as basis of vector space of all the right invariant elements of ${\cal M}$ the coordinates $P_{N}$ of the energy-momentum four-vector. $P_{0}$ and $P_{n}$ are identified to the energy operator and the coordinates of the vector-momentum operator respectively. In this case, we have only to replace in this section $X_{N}$ by $P_{N}$, $\tau^{2}$ by $G^{00}m^{2}$, the rest state $|{\cal P}_{0}\rangle$ by $|m,0,m^{2}\rangle$ and the $|{\cal P}'\rangle$ by $|E,p_{3},m^{2}\rangle$ where $E$ and $p_{3}$ are real eigenvalues of $P'_{0}$ and $P'_{3}$ respectively and $m^{2}$ is the square of rest mass which is bi-invariant, real and central. The mass is given by $E^{2}-|p|_{q}^{2}=m^{2}$ where $|p|_{q}^{2}=-\frac{G^{ij}}{G^{00}}P_{i}P_{j}$ is the length of the vector-momentum. As for the space-time coordinates we can measure exactly and simultaneously the mass, the energy, the length of the vector-momentum and only one of its components, for instance $p_{3}$. From momentum version of (30) we may see that the eigenvalues $E$ and $p_{3}$ are given in terms of the mass and the velocity by \begin{eqnarray} E=\frac{m}{(1-|v|_{q}^{2})^{\frac{1}{2}}}=m\gamma~~,~~ p_{3}=\frac{mv_{3}}{(1-|v|_{q}^{2})^{\frac{1}{2}}}=mv_{3}\gamma~~and~~ |p|_{q}=\frac{m|v|_{q}}{(1-|v|_{q}^{2})^{\frac{1}{2}}} \end{eqnarray} which are the quantum analog of the usual relativistic formulas of the energy and the vector-momentum of free particles moving in the Minkowski space with velocity $v$. \section{Evolution of particles in quantum Minkowski space-time} Having associated the quantum mechanics principles with the properties of the quantum coordinates of the Minkowski space-time and their transformations under the quantum Lorentz group, we investigated in this section the states describing the evolution of particles in the quantum the Minkowski space-time. Let $|t,x_{3},\tau^{2}\rangle$ be a state satisfying (21). From (10-11), we see that the states $|n,t,x_{3},\tau^{2}\rangle$ and $|-m,t,x_{3},\tau^{2}\rangle$ given by \begin{eqnarray} |n,t,x_{3},\tau^{2}\rangle=\overbrace{Z...Z}^{n}|t,x_{3},\tau^{2}\rangle ~and~ |-m,t,x_{3},\tau^{2}\rangle=\overbrace{\overline{Z}...\overline{Z}}^{m} |t,x_{3},\tau^{2}\rangle \end{eqnarray} are eigenstates of $X_{3}$ with eingenvalues given respectively by \begin{eqnarray} X_{3}|n,t,x_{3},\tau^{2}\rangle = (q^{2n}x_{3} + q^{-1}(q^{2n}-1)t) |n,t,x_{3},\tau^{2}\rangle=x_{3}^{(n)}|n,t,x_{3},\tau^{2}\rangle,\\ X_{3}|-m,t,x_{3},\tau^{2}\rangle = (q^{-2m}x_{3} + q^{-1}(q^{-2m}-1)t) |-m,t,x_{3},\tau^{2}\rangle=x_{3}^{(-m)}|-m,t,x_{3},\tau^{2}\rangle. \end{eqnarray} In following we shall assume that $q>0$. Therefore we can see that these eigenvalues satisfy \begin{eqnarray} x_{3}^{(n)}\leq x_{3} ~and~ x_{3}^{(-n)}\geq x_{3}~~~~~for~0<q<1,\\ x_{3}^{(n)}\geq x_{3} ~and~ x_{3}^{(-n)}\leq x_{3}~~~~~~~~for~q>1 \end{eqnarray} then for $0 <q< 1~(q > 1)$, $Z$ and $\overline{Z}$ are lowering (raising) and raising (lowering) operators for the engenvalues of $X_{3}$. To establish the conditions on the eigenvalues $x_{3}$, we consider the relations \begin{eqnarray} (a)~~Z\overline{Z}= q^{-2}(X_{0}+qX_{3})(X_{0}-q^{-1}X_{3})- q^{-\frac{1}{2}}\tau^{2},\nonumber\\ (b)~~~~\overline{Z}Z= q^{-2}(X_{0}+qX_{3})(X_{0}-q^{3}X_{3})-q^{-\frac{1}{2}}\tau^{2} \end{eqnarray} obtained from (13) and the commutation relation (12). Since the states must be of positive norm, the main value of (39) satify: \begin{eqnarray} (a)~~q^{-2}(t+qx_{3})(t-q^{-1}x_{3})-q^{-2}\alpha^{2}\geq 0,\nonumber\\ (b)~~~~q^{-2}(t+qx_{3})(t-q^{3}x_{3})-q^{-2}\alpha^{2}\geq 0 \end{eqnarray} where $\tau^{2}=q^{-\frac{3}{2}}\alpha^{2}$. From (28), we can replace $X_{i}$ by $V_{i}t$ into the norm of the vector $X_{I}$ to get $q^{-\frac{3}{2}}t^{2} - G^{ij}V_{i}V_{j}t^{2} =q^{-\frac{3}{2}}\alpha^{2}$ or $t^{2}(1-|v|_{q}^{2})=\alpha^{2}$, hence $\alpha^{2}=\frac{t^{2}}{\gamma^{2}}$. The conditions (40$a$-$b$) define the interval of the eigenvalues $x_{3}$ in which it is possible to construct states of positive norm belonging to Hilbert space. To define this interval, we begin to give the roots of (40$a$-$b$) \begin{eqnarray} x_{03}^{(a)\pm} =\frac{(q-q^{-1})\pm (Q^{2}-\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2}t~~~~~~~~~where~~ x_{03}^{(a)-}<x_{03}^{(a)+},\\ x_{03}^{(b)\pm} =-q^{-2}\frac{(q-q^{-1})\pm (Q^{2}-\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2}t~~~~~~where~~ x_{03}^{(b)+}<x_{03}^{(b)-}. \end{eqnarray} The reality of $x_{3}$ requires $Q^{2}-\frac{4}{\gamma^{2}}\geq 0$, hence $\gamma^{2}\geq\frac{4}{Q^{2}}$. It is not very difficult to see that $x_{03}^{(a)-}<x_{03}^{(b)+}$ and $x_{03}^{(a)+}<x_{03}^{(b)-}$ if $0<q< 1$ and $x_{03}^{(b)+}<x_{03}^{(a)-}$ and $x_{03}^{(b)-}<x_{03}^{(a)+}$ if $q > 1$. So the conditions (40$a$-$b$) are both satisfied in the interval $(x_{03}^{(a)-}<x_{03}^{(a)+})\cap (x_{03}^{(b)+}<x_{03}^{(b)-})$ which must be nonempty. This requires that \begin{eqnarray} (a')~~~x_{03}^{(b)+}<x_{03}^{(a)+}~~~for~0<q< 1~and~~(b')~~ x_{03}^{(a)-}<x_{03}^{(b)-}~~~for~q > 1. \end{eqnarray} The relation (43$b'$) is satisfyied if $(q-q^{-1}) - (Q^{2}-\frac{4}{\gamma^{2}})^{\frac{1}{2}}\leq - q^{-2}(q-q^{-1}) + q^{-2}(Q^{2}-\frac{4}{\gamma^{2}})^{\frac{1}{2}} \Rightarrow (1-q^{-2})Q\leq (1+q^{-2})(Q^{2}-\frac{4}{\gamma^{2}})^{\frac{1}{2}}$ or $(1-q{-2})^{2}Q^{2}\leq(1+q^{-2})^{2}(Q^{2}-\frac{4}{\gamma^{2}})$ implying \begin{eqnarray} \gamma^{2} \geq \frac{q^{2}(1+q^{-2})^{2}}{Q^{2}} = 1 \end{eqnarray} the relation (43$a'$) gives the same condition (44) on $\gamma$. This relation is necessary for the existence of Hilbert states of positive norm (physical states), hence the causality principle in noncommutative special relativity which states that there exist physical states only for real velocities in the range $0\leq |v|_{q} \leq c = 1$.\\ Now, suppose that $x_{3} \in (x_{03}^{(b)+},x_{03}^{(a)+})$, $(0<q<1)$, is an eigenvalue of the eigenstate $|t,x_{3},\tau^{2}\rangle$ of $X_{3}$. If we consider the state $|-m,t,x_{3},\tau^{2}\rangle$ eigenvector of $X_{3}$ with eigenvalue $q^{-2m}x_{3} + q^{-1}(q^{-2m}-1)t > x_{3}$ such that $q^{-2(m+1)}x_{3} + q^{-1}(q^{-2(m+1)}-1)t > x_{03}^{(a)+}$, since the state norms are positive, then \begin{eqnarray*} q^{-2m}x_{3} + q^{-1}(q^{-2m}-1)t = x_{03}^{(a)+} = \frac{(q-q^{-1})}{2}t + \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2}t \end{eqnarray*} implying \begin{eqnarray} x_{3} = q^{2m}(\frac{Q}{2} + \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2})t -q^{-1}t. \end{eqnarray} Since the operator $Z$ decreases the value of $x_{3}$, it exists a number $r\in N^{+}$ such that $\overbrace{Z...Z}^{r} |t,x_{3},\tau^{2}\rangle$ is eigenstate of $X_{3}$ with eigenvalue $x_{03}^{(b)+}$, hence \begin{eqnarray} q^{2r}x_{3} + q^{-1}(q^{2r} -1)t= -q^{-2}(\frac{(q-q^{-1})}{2}+ \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2})t. \end{eqnarray} By setting (45) into (46), we get \begin{eqnarray} q^{2(r+m)}(\frac{Q}{2} + \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2}) =q^{-2}(\frac{Q}{2} -\frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2}) \end{eqnarray} from which we obtain \begin{eqnarray*} \frac{Q^{2}}{4}(q^{2(r+m+1)} - 1)^{2} = \frac{Q^{2} -\frac{4}{\gamma^{2}}}{4}(q^{2(P+m+1)} + 1)^{2} \end{eqnarray*} leading to \begin{eqnarray} \gamma^{2} = \frac{(q^{(r+m+1)} +q^{-(r+m+1)})^{2}}{Q^{2}}. \end{eqnarray} On the other hand, if we consider the state $|n,t,x_{3},\tau^{2}\rangle$ eigenvector of $X_{3}$ with eigenvalue $q^{2n}x_{3} + q^{-1}(q^{2n}-1)t < x_{3}$ such that $q^{2(n+1)}x_{3} + q^{-1}(q^{2(n+1)}-1)t < x_{03}^{(b)-}$, we have necessary \begin{eqnarray*} q^{2n}x_{3} + q^{-1}(q^{2n}-1)t = x_{03}^{(b)-} =-q^{-2} (\frac{(q-q^{-1})}{2} + \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2})t \end{eqnarray*} implying \begin{eqnarray} x_{3}=q^{-2(n+1)}(\frac{Q}{2} - \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2})t - q^{-1}t. \end{eqnarray} Since the operator $\overline{Z}$ increases the value of $x_{3}$, there exists a number $s\in N^{+}$ such that $\overbrace{\overline{Z}...\overline{Z}}^{s} |t,x_{3},\tau^{2}\rangle$ is a eigenstate of $X_{3}$ with eigenvalue $x_{03}^{(a)+}$ yielding \begin{eqnarray} q^{-2s}x_{3} + q^{-1}(q^{-2s} -1)t= \frac{(q-q^{-1})}{2}t+ \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2}t. \end{eqnarray} By setting (49) into (50), we get \begin{eqnarray*} q^{-2(s+n+1)}(\frac{Q}{2} - \frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2}) =\frac{Q}{2} +\frac{(Q^{2} -\frac{4}{\gamma^{2}})^{\frac{1}{2}}}{2} \end{eqnarray*} which is equivalent to (47). Therefore $\gamma^{2}$ is quantized and given by \begin{eqnarray} (\gamma^{(L)})^{2} = \frac{(q^{(s+n+1)} +q^{-(s+n+1)})^{2}}{Q^{2}} = \frac{(q^{(r+m+1)} +q^{-(r+m+1)})^{2}}{Q^{2}} = \frac{(q^{(L+1)} +q^{-(L+1)})^{2}}{Q^{2}} \end{eqnarray} where $L=0,1,2,..., \infty$. By substituting (51) into (45), we get the states $|t,x_{3}^{(L,n)},\tau^{2}\rangle$ such that $X_{3}|t,x_{3}^{(L,n)},\tau^{2}\rangle =x_{3}^{(L,n)} |t,x_{3}^{(L,n)},\tau^{2}\rangle$ where \begin{eqnarray} x_{3}^{(L,n)}= (Q\frac{q^{-(L+1-2n)}}{q^{(L+1)}+q^{-(L+1)}} - q^{-1})t \end{eqnarray} and $n=0,1, ...,L$. Note that if we set $L=2l$ and $m=l-n$ we can rewrite the states as $|t,x_{3}^{(l,m)},\tau^{2}\rangle$ eigenstates of $X_{3}$ with eigenvalues \begin{eqnarray} x_{3}^{(l,m)}=q^{-1}(Q\frac{q^{2m}}{q^{(2l+1)}+q^{-(2l+1)}} - 1)t = q^{-1}(\frac{q^{2m}}{\gamma^{(l)}} - 1)t \end{eqnarray} where $\gamma^{(l)} = \frac{q^{(2l+1)} +q^{-(2l+1)}}{Q}$, with $l=0,\frac{1}{2},1,....\infty$ and $m$ runs by integer steps over the range $-l\leq m \leq l$. The substitution of (51) into (49) gives the same states with the same eigenvalues (52). Following the same analysis as above, we can show that the case $q > 1$ gives the same results for $\gamma$ (51) and for $x_{3}$ (52). From (30) and (53) we deduce the quantization of the velocity as \begin{eqnarray} |v|_{q}^{2}=1-\frac{1}{(\gamma^{(l)})^{2}}~~~and~~~ v_{3}^{(l,m)}=q^{-1}(\frac{q^{2m}}{\gamma^{(l)}} - 1). \end{eqnarray} Note that in the case where we consider the energy-momentum four-vector $P_{N}$ instead of the coordinates $X_{N}$. The relation (33) combined with (51) and (54) show that the energy $E$ and the component $P_{3}$ of the vector-momentum present discrete spectrums given by $E^{(l)}=m\gamma^{(l)}$ and $P_{3}^{(l,m)}=mv_{3}^{(l,m)}\gamma^{(l)}$. The basis of the Hilbert space states of free particle in the momentum representation is given by $|E^{(l)},p_{3}^{(l,n)},m\rangle$ normalized as $\langle E^{(l)},p_{3}^{(l,n)},m|E^{(l')},p_{3}^{(l',n')},m\rangle = \delta_{l,l'}\delta_{n,n'}$. where $l,l'=0,\frac{1}{2},1,....\infty$ and $-l\leq m,m' \leq l$. \section{Discussions and conclusions} In this paper, we have seen that an adequat association of the quantum mechanics principles with the properties of the quantum Minkowski space-time and its quantum Lorentz transformations make possible the description of the evolution of particles in the noncommutative special relativity. In this formalism we have showed that only the length of the velocity and one of its components can be measured exactly and present discret spectrums. From the quantum boost, we have established the principle of causality and the $q$-deformed analog of the lifetime dilatation formula for moving unstable particles from which we see that $\frac{\Delta(t'_{q})}{\Delta(t_{q})} =\gamma^{(l)}$, ($l,=0,\frac{1}{2},1,....\infty$) is quantized. This quantization presents an interesting novelty because as opposite to the costumary believe where we consider that the effects of noncommutativity of space may be observed only at very high energy, Planck scale $M_{P}$, which is beyond the reach of conceivable experiments, in this scenario, the effects of the evolution of free particles in the noncommutative special relativity are expressed in terms of quantized velocities (hence quantized currents for charged particles) which require not very energies but high precision on measurements.\\ \vskip0.1truecm Note that the case $L=0,~n=0$ corresponds to $x_{3}=0$ and $\gamma_{q}^{2} =1$ yielding $|v|=0$ and corresponds to the unique state describing a particle at rest considered in the section 3.\\ \vskip0.1truecm In the case where $\tau^{2}=0$, corresponding to a particle on the light cone, the roots of (40$a$-$b$) reduce to $x_{03}^{(a)-}=-q^{-1}t<x_{03}^{(a)+}=qt$ and $x_{03}^{(b)+}=-q^{-1}t<x_{03}^{(b)-}=q^{-3}t$. Then the conditions (40$a$-$b$) are both satisfied in the interval $(x_{03}^{(a)-}=x_{03}^{(b)+},x_{03}^{(a)+})$ if $0<q<1$ or in the interval ($x_{03}^{(a)-}=x_{03}^{(b)+},x_{03}^{(b)-})$ if $q >1$.The states of positive norm are given by $|t,x_{3}^{(n)},\tau^{2}\rangle$ where the engenvalues of $X_{3}$ are given by \begin{eqnarray} x_{3}^{(n)} = q^{2n}Qt-q^{-1}t,~~~~~~for~0<q\leq 1\\ x_{3}^{(n)} = q^{-2(n+1)}Qt-q^{-1}t,~~~~~for~q\geq 1 \end{eqnarray} with $n=0,1,..., \infty$. Note that the eigenstate $|t,-q^{-1}t,0\rangle$ is stable under the action of $Z$ and $\overline{Z}$ in the sens that $\forall n,m\in N^{+},~ \overbrace{Z...Z}^{n}\overbrace{\overline{Z}...\overline{Z}}^{m} |t,-q^{-1}t,0\rangle$ is an eigenstate of $X_{3}$ with eigenvalue $-q^{-1}t$ hence $v_{3}=-q^{-1}$. $\langle t,-q^{-1}t,0|Z\overline{Z}|t,-q^{-1}t,0\rangle$ and $\langle t,-q^{-1}t,0|\overline{Z} Z|t,-q^{-1}t,0\rangle $ vanish, therefore, the light velocity reduce to the $|v|_{q}^{2}=q^{2}v_{3}^{2}=1$ which is, by virtue of (31), the upper limit of velocities. We may retrieve this case in the limit $L \rightarrow \infty\Rightarrow \tau^{2} \rightarrow 0$.\\ \vskip0.1truecm To boost again the state $|t,x_{3},\tau^{2}\rangle$ we must know explicitly all the commutation relations of the sixteen generators $\Lambda_{N}^{~M}$ to get the set of commuting generators and their eigenstate $|sym_{q}\rangle$ labelled by its different eigenvalues. This study permits to define the addition rules of velocities in the noncommutative special relativity [8]. \vskip0.1truecm When we restrict the generators of the quantum $SL(2,C)$ group to those of the $SU(2)$ by imposing unitarity condition, we get relations (5) and (6) which lead us to the restriction of the Minkowski space-time transformations to the orthogonal transformations of the three dimensional space $R_{3}$ equipped with the coordinate system $X_{i}$, ($i=1,2,3$). These transformations leave invariant the time coordinate $X_{0}$. In fact, as a consequence of (5) and (6) we have \begin{eqnarray} \Delta_{L}(X_{0}) = \overline{\Lambda}_{0}^{~0}\otimes X_{0} = I\otimes X_{0}\nonumber\\ \Delta_{L}(X_{i}) = \overline{\Lambda}_{i}^{~j}\otimes X_{j} \end{eqnarray} where the generators $\overline{\Lambda}_{i}^{~j}= \frac{1}{Q} \overline{\sigma}_{i\dot{\gamma}}^{~~\alpha}M_{\alpha}^{~\sigma} \sigma^{j}_{~\sigma\dot{\rho}}M_{\dot{\beta}}^{~\dot{\rho}} \varepsilon^{\dot{\gamma}\dot{\beta}}=\frac{1}{Q} \overline{\sigma}_{i\dot{\gamma}}^{~~\alpha}M_{\alpha}^{~\sigma} \sigma^{j}_{~\sigma\dot{\rho}}S(M_{\rho}^{~\beta}) \varepsilon^{\dot{\gamma}\dot{\beta}}$ generate a Hopf sub-algebra ${\cal SO}_{q}(3)$ of $\cal L$ whose the axiomatic structure is derived from those of $\cal L$ as $\Delta(\overline{\Lambda}_{i}^{~j}) = \overline{\Lambda}_{i}^{~k}\otimes \overline{\Lambda}_{k}^{~j}$, $\varepsilon(\overline{\Lambda}_{i}^{~j})= \delta_{i}^{~j}$ and $S(\overline{\Lambda}_{i}^{~j}) = G_{iK} \overline{\Lambda}_{L}^{~K}G^{~Lj} = G_{ik} \overline{\Lambda}_{l}^{~k}G^{~lj}$ where $G^{~ij}$ is the restriction of $G^{~IJ}$ to the quantum space $R_{3}$ satisfying $G^{~ik}G_{kj} = \delta_{j}^{i} =G_{jk}G^{~ki}$. The form of the antipode of $\overline{\Lambda}_{i}^{~j}$ implies the orthogonality properties \begin{eqnarray*} G^{~ij} \overline{\Lambda}_{i}^{~l}\overline{\Lambda}_{j}^{~k} = G^{~lk} &~and& G_{lk} \overline{\Lambda}_{i}^{~l} \overline{\Lambda}_{i}^{~k}= G_{ij}. \end{eqnarray*} Therefore, $\overline{\Lambda}_{i}^{~j}=\frac{1}{Q} \overline{\sigma}_{i\dot{\gamma}}^{~~\alpha}M_{\alpha}^{~\sigma} \sigma^{j}_{~\sigma\dot{\rho}}S(M_{\rho}^{~\beta}) \varepsilon^{\dot{\gamma}\dot{\beta}}$ establishes a correspondence between $SU_{q}(2)$ and $SO_{q}(3)$ group. In fact the two-dimensional representation of $SU_{q}(2)$ is given by $\left(\begin{array}{cc} \alpha& q\gamma^{\star}\\-\gamma&\alpha^{\star} \end{array} \right)$[9] and in the three dimensional space spanned by the basis $Qe_{-1}=Z$, $e_{0} =X_{3}$ and $Qe_{1}=\overline{Z}$, the generators $\overline{\Lambda}_{i}^{~j}=\frac{1}{Q} \overline{\sigma}_{i\dot{\gamma}}^{~~\alpha}M_{\alpha}^{~\sigma} \sigma^{j}_{~\sigma\dot{\rho}}S(M_{\rho}^{~\beta}) \varepsilon^{\dot{\gamma}\dot{\beta}}$ write \begin{eqnarray*} (d_{1,ij})_{i,j=-1,0,1}= \left(\begin{array}{clcr} \alpha^{\star 2} & -(1+q^{2})\alpha^{\star}\gamma &-q\gamma^{2}\\ \gamma^{\star}\alpha^{\star} & 1-(1+q^{2})\gamma^{\star}\gamma & \alpha\gamma\\ -q\gamma^{\star 2}& -(1+q^{2})\gamma^{\star}\alpha & \alpha^{2} \end{array} \right)\in M_{3}\otimes C(SU_{q}(2)) \end{eqnarray*} which is an irreducible three-dimensional representation of $SU_{q}(2)$ considered in [9]. In the other hand if we replace $\overline{Z}=Qe_{1}$, $X_{3}=e_{0}$, $Z=Qe_{-1}$, $\lambda = (q-q^{-1})t=(q-q^{-1})X_{0}$ and $\rho =q^{-2}t^{2} - q^{-\frac{1}{2}}\tau^{2}$ into (19-22) we retrieve the algebra of the quantum spheres given by (2$a$-$e$) of [10]. In the framework developped above, the states $|t,x_{3}^{(n)},\tau^{2}\rangle$ whose eigenvalues are given by (55-56) (case $\tau^{2}=0$) represent a basis of the Hilbert space where the quantum sphere algebra is represented. The fixed time $t$ represents the ray of the sphere.\\ \vskip0.1truecm {\bf Acknowledgments.} I am grateful to M. Dubois-Violette for his kind interest and hepful suggestions. I am grateful for hospitatility at the Abdus Salam International centre for theoretical physics where the extended version of this work was done.\\ {\bf References:} 1)U. Meyer, Commun. Math. Phys. 174(1996)457. P. Podle\'s, Commun. Math. Phys. 181(1996)569.\\ 2)M. Pillin, W. B. Schmidke and J. Wess, Nucl. Phys. B(1993)223. M. Pillin and Weilk, J. Phys. A: Math. Gen. 27(1994)5525. B. L. Cerchiai and J. Wess :"q-Deformed Minkowski Space based on q-Lorentz Algebra" LMU-TPW 98-02, MPI-PhT/89-09.\\ 3)M. Lagraa, "On the quantum Lorentz group", to appear in J. Geom. Phys.\\ 4)M. Lagraa, Int. J. Mod. Phys. A11(1996)699.\\ 5)M. Dubois-Violette and G. Launer, Phys. Lett. B245(1990)175.\\ 6)O. Ogiesvestsky, W. B. Schmidke, J. Wess and B. Zumino, Commun. Math. Phys. 150(1992)495.\\ 7)U. Carow-Watamura, M. Schlieker, M. Scholl and S. Watamura, Z. Phys. C- $particles~and~fields$ 48(1990)159, Int. J. Mod. Phys. A6(1991)3081.\\ 8)M. Lagraa, work in progress.\\ 9)S. L. Woronowicz, Publ. RIMS, Kyoto Univ. 23 (1987) 117, Commun. Math. Phys. 122(1989)125.\\ 10)P. Podle\'s, Lett. Math. Phys. 14(1987)193, Lett. Math. Phys. 18(1989)107, Commun. Math. Phys. 170(1995)1.\\ \end{document} \baselineskip=24pt 6)M. Pillin, J. Math. Phys. 35(1994)2804. B. L. Cerchiai and J. Wess :"q-Deformed Minkowski Space based on q-Lorentz Algebra"
\section{Introduction} \label{intro} \footnotetext[1]{Submitted to {\em Dynamics: Models and Kinetic Methods for Nonequilibrium Many-Body Systems}, Ed. J. Karkheck, Kluwer, Dordrecht, The Netherlands, Feb. 1999.} In rapid flows of granular media, the mean time between collisions of grains is much longer than the duration of a collision~\cite{campbell90}; for such flows, the machinery of kinetic theory is expected to apply. Continuum equations~\cite{lun84,jenkins85} analogous to the Navier-Stokes equations can be produced, allowing quantitative analysis of flows. The simplest and most common formulations incorporate Boltzmann's assumption of molecular chaos: that particle velocities are uncorrelated. While this assumption works well for low-density molecular gases, granular gases may not abide such a restriction because collisions between grains are inelastic. Inelastic collisions reduce relative velocities, so that post-collisional velocities are more parallel than pre-collisional velocities. Repeated inelastic collisions can lead to strong, long-range velocity correlations, which standard kinetic theory does not include. We will use molecular dynamics simulations to produce steady state granular gases and study the velocity correlations that develop. The importance and intrinsic interest of velocity correlations in granular flows have been noted by a number of researchers. Two-dimensional simulations of an initially homogeneous distribution of inelastic disks without velocity correlations show that as time progresses, velocity correlations build in both strength and range~\cite{orza98}. These simulations are limited in time, however, because the homogeneous state is unstable to density fluctuations, and rapidly becomes inhomogeneous. Nevertheless, these simulations clearly displayed a characteristic vortex structure of the correlations. Based upon similar considerations, ring kinetic theory, which accounts for velocity correlations, has been applied to the cooling state~\cite{vannoije98}. One-dimensional simulations of stochastically forced point particles also show velocity correlations~\cite{swift98}. We apply stochastic forcing~\cite{swift98,williams96} to two-dimensional event-driven simulations of inelastic disks. The forcing overcomes the tendency of the granular material to form density clusters, and approximately homogeneous steady states form. In an earlier study of these states~\cite{bizon99}, we found strong velocity correlations that extended throughout the entire simulation area. In the present work, we discuss the simulation method, show that the velocity correlations are essentially independent of the simulated area, and describe the vortex structure of the correlations. \section{Simulations of Driven Granular Gases} We treat collisions between molecules as instantaneous and binary. The collisions between grains conserve momentum but dissipate energy. Between collisions, particles travel along straight lines if unaccelerated, or along parabolas if accelerated. This model allows efficient simulation of collections of particles using event-driven molecular dynamics~\cite{lubachevsky91,marin93}. When particles collide, the component of the relative particle velocity along the line joining particle centers, $v_n$, is reversed, and reduced by a factor $e$, the coefficient of restitution, which can take values between 1 for elastic particles and 0 for completely inelastic particles. We allow $e$ to depend on $v_n$ through \begin{equation} e(v_n) = \left \{\begin{array}{cc} 1 - B v_n ^ {\beta} &, v_n < v_o \\ \epsilon &, v_n > v_o \end{array} \right. , \label{restform} \end{equation} where $B = (1-\epsilon)(v_o)^{-\beta}$, $\beta=3/4$ and $\epsilon$ is a constant, chosen to be $0.7$. These parameters give quantitative agreement to experiments on patterns in vertically oscillated granular media~\cite{bizon98,debruyn98}. The variation in $e$ has the effect of removing inelastic collapse~\cite{goldman98}, which is a singularity in the inelastic hard sphere model that produces an infinite number of collisions within a finite time~\cite{mcnamara92,mcnamara94}. In general, colliding particles also exert frictional forces on one another; for this paper, we assume that the coefficient of friction is zero, so that we are studying only the effects of inelasticity. Because of inelasticity, the energy of an unforced collection of grains inevitably decreases. To achieve steady states, then, we must force the granular material. Methods that force through boundaries, such as shaking, invariably produce strong inhomogeneities in the system; to achieve near-homogeneity, we force volumetrically, assuming the particles to be in contact with a white-noise heat bath~\cite{williams96}. Whenever two particles collide, the velocities of two other randomly selected particles are changed by amounts $|\delta {\bf v}| \hat{\bf r}_i$, where the magnitude of the kicks, $|\delta {\bf v}|$, are always the same, but the direction vectors, $\hat{\bf r}_i$ are randomly chosen for each kicked particle. In addition to the white noise heat bath, we perform a lesser number of runs with two other heat baths. To model the motions of pucks on an air table~\cite{oger96,ippolito95}, we can allow particles to accelerate randomly from collision to collision. Finally, we model the effects of a strong heat bath, which we denote the Boltzmann bath, by completely obliterating the velocities of randomly chosen particles, and giving new velocities based on a Boltzmann distribution. The details of all three forcing methods may be found in ~\cite{bizon99}. We perform simulations of N disks of diameter $\sigma$ moving in a two-dimension square of side length $L$, which varies from $52.6 \sigma$ to $420.8 \sigma$. The simulation box is periodic in both directions. The solid fraction, defined as $N {{\pi}\over{4}} {{\sigma^2}\over{L^2}}$, is $0.5$ for all runs. Because of the variation of $e$ with relative normal velocity, the velocity scale $v_0$ enters; we use $v_0$ to nondimensionalize velocities, and $v_0^2$ to nondimensionalize the granular temperature $T$. For $T$ much larger than one, most particle collisions will occur with the high-velocity value of $e$, $0.7$; for lower $T$, a range of $e$ will occur. \section{Dependence of Correlations upon Simulation Area} We denote two particles $1$ and $2$, and ${\bf \hat{k}}$ the a unit vector pointing from the center of $1$ to the center of $2$. The velocity of $1$ then has a components parallel to, $v_{1}^{||}$, and perpendicular to, $v_{1}^{\perp}$, ${\bf \hat{k}}$, as does particle $2$. We define two correlation functions \begin{eqnarray} \langle v_{1}^{||} v_{2}^{||} \rangle &=& \sum v_{1}^{||} v_{2}^{||} / N_r,\\ \langle v_{1}^{\perp} v_{2}^{\perp} \rangle &=& \sum v_{1}^{\perp} v_{2}^{\perp} / N_r, \end{eqnarray} where the sums are over the $N_r$ particles such that the distance between the two particles is within $\delta r$ of $r$. For uncorrelated particle velocities, $\langle v_{1}^{||} v_{2}^{||} \rangle$ and $\langle v_{1}^{\perp} v_{2}^{\perp} \rangle$ will both give zero. In the smallest simulation area, $L=52.6\sigma$, correlations extend the full length of the computational cell. Cell filling structures may be divided into two cases: structures with a natural length that is larger than the box in which they exist and structures that will always grow to fill any finite box. To differentiate between the former and the latter, we performed four simulations with white noise forcing, quadrupling the area at each step, while holding the solid fraction fixed at $0.5$. The granular temperature $T$ is approximately $30$, but varies between $28$ in the smallest box and $32$ in the largest. This variation in temperature is not important; for $T>>1$, the coefficient of restitution is independent of collision velocity. In this limit, the role of the temperature is simply to set the velocity scale. The velocity correlation functions are shown in Fig.~\ref{vcorrsize}. Even in the largest simulation, composed of 112768 particles, the correlations fill the box. However, the correlation functions for the largest simulation are somewhat different from the smaller ones. This is probably due to poorer statistics; in terms of collisions per particle, this run lasted only one-half as long as the next largest. \begin{figure} \epsfxsize=.95\textwidth \centerline{\epsffile{VcorrelationsL.ps}} \smallskip \caption{Velocity correlations as a function of particle separation at $\nu=0.5$ and $T \approx 30$, for four different box sizes. $+: L=52 \sigma$, $\triangle: L=105 \sigma$, $*: L=211 \sigma$, $\Box: L=421 \sigma$.} \label{vcorrsize} \end{figure} Because velocity correlations are positive for small separations, particles collide less frequently and with less relative velocity than elastic particles at the same density, for which velocity correlations are much smaller. As a result, less momentum will be transferred through inelastic collisions than through elastic collisions, and the pressure, $P$, will decrease. Assuming that velocity correlations do not exist, the equation of state for dense granular gases is given by~\cite{jenkins85} \begin{equation} P = (4/\pi\sigma^2) \nu T (1 + (1+e) G(\nu)). \label{state}, \end{equation} The first term on the right hand side, $(4/\pi\sigma^2) \nu T$, accounts for momentum transfer due to particle streaming without collisions, while the second term, $(4/\pi\sigma^2) \nu T (1+e) G(\nu)$, accounts for the momentum transfer due to particle collisions~\cite{chapman}. In the absence of velocity correlations, $G(\nu)$ is defined as $\nu g(\nu,\sigma)$, where $g(\nu,\sigma)$ is the radial distribution function for the particles, evaluated at zero particle separation. Calculation of $P$ from simulation, via measurement of the virial~\cite{rapaportsbook}, becomes a measurement of $G(\nu)$, which describes the collisional momentum transport. If velocity correlations exist, $G(\nu)$ will be reduced, since less momentum will be transported collisionally. Figure ~\ref{vcorrsize} shows that the short range velocity correlations depend on the size of the box; therefore, $G(\nu)$ should also depend on $L$. Figure~\ref{GvsL} displays $G(\nu)$ as a function of $L$ for these four runs. Over about one decade, $G(\nu)$ scales with $\log{L}$. Clearly this scaling can not continue indefinitely, since unphysical negative values of $G(\nu)$ would result. Note also, that increasing the box size actually leads to values of $G(\nu)$ farther from the values for uncorrelated velocities. \begin{figure} \epsfxsize=.9\textwidth \centerline{\epsffile{GofL.ps}} \smallskip \caption{$G$ as a function of $L$ for the runs shown in Fig.~\protect{\ref{vcorrsize}}. $L_o=52\sigma$ denotes the length of the smallest box. The dotted line is a fit to all four points: $G(\nu) = 1.3 - 0.04 \log_2(L/L_o)$, The solid line as a fit to the three largest $L$ values: $G(\nu) = 1.295 - 0.038 \log_2(L/L_o)$. Note that the log is base 2.} \label{GvsL} \end{figure} \begin{figure} \epsfxsize=.9\textwidth \centerline{\epsffile{PvcL.ps}} \smallskip \caption{Probability distribution of collision velocities $v_c = |{\bf v}_1 - {\bf v}_2|$, for the data in Figs.~\protect{\ref{vcorrsize}} and~\protect{\ref{GvsL}}. $+: L=52 \sigma$, $\triangle: L=105 \sigma$, $*: L=211 \sigma$, $\Box: L=421 \sigma$. The solid curve is $P(v_c/\sqrt{T}) = (1/2\sqrt{\pi T^3}) v_c^2 e^{-v_c^2/4T}$, which holds for elastic particles.} \label{PvsL} \end{figure} This unusual result, that the importance of velocity correlations increases with increasing computational area, can also be deduced from the distribution of collision velocities. Figure~\ref{PvsL} exhibits these distributions for the runs displayed in Figures~\ref{vcorrsize} and~\ref{GvsL}. As the computational area increases, so too does the deviation from the distribution predicted for particles chosen without correlation from a Boltzmann distribution, plotted as a solid curve. \section{Vortex Structure} Inelasticity breeds velocity correlations; reduction of relative velocity in collisions leads to particles moving more alike after collisions than before. On average, then, particles will be surrounded by particles that are moving along with them. The structure of the velocity correlations can be elucidated by calculating this average flow around each particle. For a single particle $i$, we can calculate the flow around it by translating it to the origin, and rotating so that its velocity lies along the positive $x$ axis. If ${\bf v}(x,y)$ is the velocity field defined by the particles, then the flow around particle $i$ is given by \begin{equation} {\bf u}_i = R_{\theta(i)} {\bf v}(x-x_i,y-y_i), \label{floweqn} \end{equation} where $\theta(i)$ is the angle between the $i$-th particle velocity, ${\bf v}_i$ , and the positive $x$ axis, $(x_i,y_i)$ is the position of the $i$-th particle, and $R_\theta$ is the operator that rotates vectors clockwise through angle $\theta$. The average flow around particles, then, is \begin{equation} {\bf u} = \sum_{i=1}^N {\bf u}_i / N. \end{equation} Finally, ${\bf u}$ is averaged over about 100 frames to reduce noise. Figure ~\ref{Flow} displays vector fields of the average flow around particles, ${\bf u}$, for the three types of forcing, as well as for unforced elastic particles, all at $\nu=0.5$ and $T=1.05$. In each case, the vector at the origin, which measures only the average particle speed, has been suppressed, and the longest remaining vector in each field has been scaled to unit length. In both the white noise and accelerated forcings, the average flow near the origin is along the positive $x$ axis, {\it i.e.,~} with the direction of the central particle's motion. The Boltzmann bath shows some indications of this effect close to the origin, but the correlations are destroyed by the strongly thermalizing forcing before they can propagate to larger length scale. For the elastic particles, there is no discernible flow, only noise. \begin{figure} \epsfxsize=.9\textwidth \centerline{\epsffile{FieldsT0.ps}} \smallskip \caption{The average velocity fields around a particle centered in each cell and moving to the right, ${\bf u}$, for elastic particles and for inelastic particles forced in three different ways (cf section 2). Each vector field is scaled separately so that its longest vector has length one. Compared to the (suppressed) central vector, these lengths are: White noise, 0.2; Accelerated, 0.27; Boltzmann, 0.008; Elastic 0.008. The boxed regions in the white noise and elastic flows are shown in Fig.~\protect{\ref{closeup}}.} \label{Flow} \end{figure} Close to any particle, surrounding particles move along with it. Farther away, the correlations decay and cannot be seen on Fig~\ref{Flow}, so the boxed regions for the white noise forcing and for elastic particles are expanded in Fig.~\ref{closeup}. While expansion of the velocity field for elastic particles produces still more noise, the inelastic flow field reveals a highly ordered vortex structure. Along the direction of the central particle's motion, the velocities slowly drop to zero, while perpendicular to the original particle's motion, the velocities drop to zero and increase in the negative direction; this flow makes clear the structure of the velocity correlation functions in Fig.~\ref{vcorrsize}. \begin{figure} \epsfxsize=.9\textwidth \centerline{\epsffile{CloseFieldsT0.ps}} \smallskip \caption{A close-up on the boxed regions in Fig~\protect{\ref{Flow}} reveals that for inelastic particles, large vortices form, one on each side of the particle. The longest vector in the velocity field for inelastic particles represents a velocity nine times larger than that represented by the longest vector for elastic particles.} \label{closeup} \end{figure} This vortical flow is reminiscent of similar structures produced in simulations of elastic particles~\cite{alder68,alder70a} by Alder and Wainwright. In their simulations, they discovered diffusive behavior different from that predicted by kinetic theory. The diffusion constant may be written in terms of the slope of the exponentially decaying autocorrelation function. However, Alder and Wainwright found deviations from exponential decay, and traced the deviations to a vortical flow. If particles $a$ and $b$ are initially uncorrelated, an elastic collision will correlate each particle's post collision velocity with the other particle's pre-collision velocity; both particles now have a correlation with the original velocity of particle $a$. As particle $b$ collides with other particles, they gain information about particle $a$'s initial velocity. Several collision times later, this information has been transmitted to many particles. There are two main differences between the vortices in flows of elastic particles and those in flows of inelastic particles. Alder and Wainwright produced the flow field given by \begin{equation} {\bf u}(t')_i = R_{\theta(i,t)} {\bf v}(x-x_i(t),y-y_i(t),t'). \end{equation} For $t'=t$, Eq.~\ref{floweqn} is recovered; for elastic particles, no structure is apparent. It is only at later times, $t' > t$, that a vortex appears in ${\bf u}(t')$. For the inelastic particles, however, structure is clear at $t'=t$. The second difference is the strength of the vortex. The strongest velocity in Alder and Wainwright's vortex was about 2\% of the original velocity, while for inelastic particles, the strongest velocity can be about 40\% of the central velocity. \begin{figure} \epsfxsize=.9\textwidth \centerline{\epsffile{InEl.ps}} \smallskip \caption{Snapshots of simulations with white noise driving and with elastic particles; large coherent structures are visible for the dissipative system on the left. Particles with positive horizontal velocity are black, particles with negative horizontal velocity are white. ($\nu = 0.5, T=1.05$)} \label{ParticleFields} \end{figure} The inelastic vortex is so strong that hints of it are visible even in a single snapshot of particles. Figure~\ref{ParticleFields} shows such a snapshot, with particles colored black if they have positive horizontal velocity and white if they have a negative horizontal velocity. For elastic particles, the black and white are well mixed, but in the inelastic case larger scale structure can be glimpsed. Black particles are concentrated along the top and bottom of the image, and white particles are concentrated along the central region. \section{Conclusion} The correlations we have found are consistent with those of simulations on the homogeneous cooling state~\cite{orza98}. In those simulations, the range of velocity correlations grew until the onset of large scale density variations. The addition of forcing in our simulations suppresses the growth of density fluctuations, allowing the velocity correlations to continue to grow until they extend throughout the entire computational area. The results we obtain are not particularly sensitive to the exact form of the forcing. In both the white noise and accelerated forcing schemes, vortical correlation structures form. Only when the bath explicitly destroys correlations, as in the Boltzmann bath, do the results differ. \section{Acknowledgments} We thank J. T. Jenkins, M. H. Ernst, and T. P. C. van Noije for useful discussions. This work was supported by the Department of Energy Office of Basic Energy Sciences.
\section{Introduction} \label{intro} Attempting to assign any sort of quantitative ranking scheme, as a measure of an astronomy graduate program's ``quality'', is a thankless task, and one almost certainly bound to be met with controversy, if not outright derision. Regardless, such rankings are, in fact, published on a semi-regular basis, most noticeably as part of the Gourman (1997) and U.S. News Graduate School Rankings ({\tt http://www.usnews.com/usnews/edu/bey-\break ond/}). In both cases, though, the rankings are effectively a subjective measure of the astronomical community's perception of each school's graduate program.\footnote{Such subjective biases are, unfortunately, a reality students should be aware of, although we do not wish to dwell upon them here.} Very few attempts at objectively quantifying the quality of a given astronomy graduate program exist, the two most notable being Domen \& Thronson (1988) and Trimble (1991). In the former, Domen \& Thronson conclude that, in the mean, graduates of Berkeley, Harvard, Caltech, Princeton, and Chicago, were more than six times as successful as graduates of Arizona, UCLA, Colorado, Minnesota, and Virginia,\footnote{The first versus bottom five entrants of their Table II.} in obtaining junior\footnote{By ``junior'', we mean associate and assistant professorships.} faculty positions at the 32 major institutes comprising their survey sample. In the latter, Trimble concludes that while $\sim 80$\% of graduates from one ``prestigious'' university are typically still involved in astronomy research, the fraction from a comparison ``non-prestigious'' program is only $\sim 50$\%.\footnote{These percentages are based upon Trimble's (1991) Table 1, combining the entries for graduates employed at PhD-granting \it and \rm government (or industrial) labs, along with the small fraction of graduates working in support duties (both hardware and software) who still maintain a modest publication record. The resulting sum should (roughly) parallel our selection criteria, as described in Section \ref{method}.} Both the Domen \& Thronson (1988) and Trimble (1991) studies provide crucial quantitative evidence in support of the hypothesis of a hierarchy of quality in astronomy graduate programs, although our contention is that there are some subtle effects within the numbers which suggest the situation is perhaps not as grim as it would appear on the surface. For example, Domen \& Thronson only consider the PhD source for professorial employees of the 32 large degree-granting institutes in the USA - i.e., the sample is biased, to some degree. One is left wondering about those graduates employed at the numerous small colleges and non-degree-granting universities, along with those in industry and government laboratories, who still remain active members of the astronomy research community. Conversely, Trimble's comparison is based upon only two universities, and one might ask if these two are truly representative of the community at large. Addressing the (potential) shortcomings of these earlier studies is the focus of our current analysis. In Section \ref{method}, we describe the methodology employed in determining the present-day (circa December 1998) status of 897 1975-1994 astronomy PhD graduates from 19 different schools in 4 different countries. While clearly not intended to be 100\% complete\footnote{Although, our sample does represent $\sim 1/4$ of the astronomy PhDs granted during the 20-year 1975-1994 baseline.}, the sample is fairly representative and unbiased, including examples of those schools traditionally considered as prestigious, as well as lesser-appreciated large and small programs, from several countries. Our putative ranking scheme is a simple objective one, based solely upon the fraction of a given school's PhD graduates who are still involved in astronomical research today, regardless of the perceived status of a given researcher's present-day institution. More complicated schemes could be envisioned whereby additional weight is ascribed to, say, publication frequency, citation history, grant application and/or observing proposal success rates, etc, but we are strongly of the opinion that \it the simplest, and perhaps ultimate, measure of a given program, is the success rate of its graduates in finding long-term astronomical research careers\rm. The main results and conclusions of our study are drawn in Section \ref{discussion}, and summarized in Section \ref{summary}. \section{Methodology} \label{method} Ideally, one would like to examine the success rates of \it all \rm the astronomy and astrophysics degree-granting institutions.\footnote{Doubly so, since it is safe to say that the first thing every reader of this paper will do is look for their own institution in Tables 1 and 2!} Unfortunately, with 4695 theses listed in NASA's Astrophysics Data System (ADS) database (for the 1975-1994 baseline of our study), of which 3700 were classified as ``astronomy/astrophysics'',\footnote{Eliminating the $\sim 20$\% of ADS-listed PhDs which, for example, are classified as particle physics, atmospheric, solar, or lunar/planetary; we should stress though that our conclusions are \it not \rm dependent upon the exclusion of these ``non-astronomy'' PhDs from the sample.}, 100\% completeness was just not feasible. Instead, representative large versus small, and prestigious versus non-prestigious, institutions were randomly selected for study. For two of the universities in our study, the ADS list of PhD graduates was compared against the respective institutes' official records, and found to be 100\% complete. For the American universities, the top three from the 1998 U.S. News Astrophysics Graduate School Rankings (Caltech, Princeton and Harvard), three from just outside the top ten (Colorado, Maryland and Hawaii), and two unranked (New Mexico State (NMSU) and Wyoming) programs were selected. These eight schools accounted for $\sim 1/7$ of the ADS-listed astronomy PhDs granted during 1975-1994. For the Canadian sample, we simply included all eight universities which granted more than ten PhDs during the same 20-year baseline. For these 16 North American institutes, their respective 1975-1994 graduate lists were culled from the ADS master list of 3700. The non-North American institutions were more problematic, as the vast majority are either not included in ADS (the norm) or woefully incomplete - as such, only three universities fall into this category, two from Australia (Mount Stromlo \& Siding Spring Observatories and Sydney University) and our sole European entrant (Leiden). For these three, we either had local access to a complete set of Annual Reports (within which year-by-year graduate information was available) or a contact at the institute in question had the relevant information in a readily available electronic form. While it would have been desirable to include, for example, several U.K. universities, year-by-year graduate information was not available to the authors. In Table 1, the 19 institutions in our study are listed in descending order of total number of astronomy PhDs granted ($n_{\rm g}$) during 1975-1994 (column 14); over an order of magnitude difference exists between the largest and smallest program. Five-year sub-samples (columns 2,5,8,11) are provided, demonstrating the overall trend of increased astronomer production, a point to which we return in Section \ref{discussion}. Again, the 897 graduates tracked in this analysis represent $\sim 1/4$ of the total number of astronomy PhDs (3700) in the ADS; our sample should be a fairly representative one of the population as a whole. \small \begin{table*} \caption{Number of astronomy and astrophysics PhD graduates ($n_{\rm g}$) per institution, along with the number still involved in astronomical research (n$_{\rm astr}$), as of late-1998, and the number with identifiably permanent or tenure-track (n$_{\rm perm}$) positions. The number of graduates of uncertain status are noted in parenthesis.} \begin{center} \begin{tabular}{lp{0.16in}p{0.35in}p{0.25in}p{0.16in}p{0.35in}p{0.25in}p{0.16in}p{0.35in}p{0.25in}p{0.16in}p{0.35in}p{0.25in}p{0.16in}p{0.35in}p{0.25in}}\hline Institution & \multicolumn{3}{c}{\hspace{-0.2in}1975-1979} & \multicolumn{3}{c}{\hspace{-0.2in}1980-1984} & \multicolumn{3}{c}{\hspace{-0.2in}1985-1989} & \multicolumn{3}{c}{\hspace{-0.2in}1990-1994} & \multicolumn{3}{c}{\hspace{-0.2in}Total} \\ & n$_{\rm g}$ & \hspace{-0.10in}n$_{\rm astr}$ & \hspace{-0.20in}n$_{\rm perm}$ & n$_{\rm g}$ & \hspace{-0.10in}n$_{\rm astr}$ & \hspace{-0.20in}n$_{\rm perm}$ & n$_{\rm g}$ & \hspace{-0.10in}n$_{\rm astr}$ & \hspace{-0.20in}n$_{\rm perm}$ & n$_{\rm g}$ & \hspace{-0.10in}n$_{\rm astr}$ & \hspace{-0.20in}n$_{\rm perm}$ & n$_{\rm g}$ & \hspace{-0.10in}n$_{\rm astr}$ & \hspace{-0.20in}n$_{\rm perm}$ \\\hline Caltech & 38 & \hspace{-0.065in}29 & \hspace{-0.175in}21(7) & 31 & \hspace{-0.065in}21 & \hspace{-0.175in}17(2) & 26 & \hspace{-0.065in}19 & \hspace{-0.175in}11(1) & 37 & \hspace{-0.065in}21 & \hspace{-0.11in}8(1) & \hspace{-0.055in}132 & \hspace{-0.065in}90 & \hspace{-0.175in}57(11)\\ Princeton & 25 & \hspace{-0.065in}18 & \hspace{-0.175in}14(2) & 20 & \hspace{-0.065in}13 & \hspace{-0.11in}7(3) & 27 & \hspace{-0.065in}22 & \hspace{-0.175in}15(1) & 29 & \hspace{-0.065in}21 & \hspace{-0.175in}10(2) & \hspace{-0.055in}101 & \hspace{-0.065in}74 & \hspace{-0.175in}46(8) \\ Harvard & 26 & \hspace{-0.065in}19 & \hspace{-0.175in}16(2) & 20 & \hspace{-0.065in}17 & \hspace{-0.175in}13(2) & 24 & \hspace{-0.065in}17 & \hspace{-0.175in}10(2) & 21 & \hspace{-0.065in}17 & \hspace{-0.11in}5(2) & 91 & \hspace{-0.065in}70 & \hspace{-0.175in}44(8)\\ Leiden & 18 & \hspace{-0.065in}11 & \hspace{-0.11in}9(2) & 22 & \hspace{-0.065in}14 & \hspace{-0.175in}13 & 26 & \hspace{-0.065in}14 & \hspace{-0.175in}10(3) & 24 & \hspace{-0.065in}14 & \hspace{-0.11in}2(2) & 90 & \hspace{-0.065in}53 & \hspace{-0.175in}34(7) \\ Maryland & 26 & \hspace{-0.065in}14 & \hspace{-0.11in}9(4) & 16 & \hspace{-0.065in}11 & \hspace{-0.11in}5(4) & 22 & \hspace{-0.065in}17 & \hspace{-0.11in}5(6) & 20 & \hspace{-0.065in}18 & \hspace{-0.11in}5(1) & 84 & \hspace{-0.065in}60 & \hspace{-0.175in}24(15)\\ MSSSO & \hspace{0.062in}9 & \hspace{-0.005in}6 & \hspace{-0.11in}5 & 18 & \hspace{-0.065in}11 & \hspace{-0.11in}6 & 16 & \hspace{-0.005in}9 & \hspace{-0.11in}6 & 25 & \hspace{-0.065in}15 & \hspace{-0.11in}3 & 68 & \hspace{-0.065in}41 & \hspace{-0.175in}20 \\ Colorado & 12 & \hspace{-0.005in}5 & \hspace{-0.11in}5 & 10 & \hspace{-0.005in}8 & \hspace{-0.11in}6(1) & 24 & \hspace{-0.065in}16(2) & \hspace{-0.11in}8(3) & 20 & \hspace{-0.065in}14(2) & \hspace{-0.11in}3(1) & 66 & \hspace{-0.065in}43(4) & \hspace{-0.175in}22(5)\\ Toronto & 15 & \hspace{-0.005in}7 & \hspace{-0.11in}6 & 18 & \hspace{-0.005in}8 & \hspace{-0.11in}6 & 17 & \hspace{-0.005in}9 & \hspace{-0.11in}8 & 11 & \hspace{-0.005in}6 & \hspace{-0.11in}1(1) & 61 & \hspace{-0.065in}30 & \hspace{-0.175in}21(1) \\ Sydney & 12 & \hspace{-0.005in}4 & \hspace{-0.11in}4 & \hspace{0.062in}3 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & 14 & \hspace{-0.065in}12 & \hspace{-0.11in}5 & 17 & \hspace{-0.005in}9(1) & \hspace{-0.11in}2 & 46 & \hspace{-0.065in}26(1) & \hspace{-0.175in}12 \\ Hawaii & \hspace{0.062in}6 & \hspace{-0.005in}5 & \hspace{-0.11in}4(1) & \hspace{0.062in}3 & \hspace{-0.005in}2 & \hspace{-0.11in}1 & \hspace{0.062in}3 & \hspace{-0.005in}3 & \hspace{-0.11in}2(1) & 14 & \hspace{-0.065in}10 & \hspace{-0.11in}1 & 26 & \hspace{-0.065in}20 & \hspace{-0.113in}8(2) \\ NMSU & \hspace{0.062in}6 & \hspace{-0.005in}3(1) & \hspace{-0.11in}2(2) & \hspace{0.062in}2 & \hspace{-0.005in}0(1) & \hspace{-0.11in}0(1) & \hspace{0.062in}9 & \hspace{-0.005in}6 & \hspace{-0.11in}3(1) & \hspace{0.062in}5 & \hspace{-0.005in}3 & \hspace{-0.11in}0(1) & 22 & \hspace{-0.065in}12(2)& \hspace{-0.113in}5(5) \\ UBC & \hspace{0.062in}8 & \hspace{-0.005in}5 & \hspace{-0.11in}4(1) & \hspace{0.062in}3 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & \hspace{0.062in}3 & \hspace{-0.005in}3 & \hspace{-0.11in}1 & \hspace{0.062in}7 & \hspace{-0.005in}4 & \hspace{-0.11in}1 & 21 & \hspace{-0.065in}13 & \hspace{-0.113in}7(1) \\ Western & \hspace{0.062in}7 & \hspace{-0.005in}2 & \hspace{-0.11in}2 & \hspace{0.062in}2 & \hspace{-0.005in}0 & \hspace{-0.11in}0 & \hspace{0.062in}4 & \hspace{-0.005in}3 & \hspace{-0.11in}1(1) & \hspace{0.062in}4 & \hspace{-0.005in}3 & \hspace{-0.11in}1 & 17 & \hspace{-0.005in}8 & \hspace{-0.113in}4(1) \\ Montreal & \hspace{0.062in}0 & \hspace{-0.005in}0 & \hspace{-0.11in}0 & \hspace{0.062in}1 & \hspace{-0.005in}0 & \hspace{-0.11in}0 & \hspace{0.062in}2 & \hspace{-0.005in}2 & \hspace{-0.11in}0(1) & 11 & \hspace{-0.005in}9 & \hspace{-0.11in}1(2) & 14 & \hspace{-0.065in}11 & \hspace{-0.113in}1(3) \\ Wyoming & \hspace{0.062in}1 & \hspace{-0.005in}0 & \hspace{-0.11in}0 & \hspace{0.062in}5 & \hspace{-0.005in}3 & \hspace{-0.11in}2(1) & \hspace{0.062in}1 & \hspace{-0.005in}1 & \hspace{-0.11in}0 & \hspace{0.062in}7 & \hspace{-0.005in}5 & \hspace{-0.11in}1(1) & 14 & \hspace{-0.005in}9 & \hspace{-0.113in}3(2) \\ Victoria & \hspace{0.062in}1 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & \hspace{0.062in}4 & \hspace{-0.005in}4 & \hspace{-0.11in}4 & \hspace{0.062in}2 & \hspace{-0.005in}2 & \hspace{-0.11in}0(1) & \hspace{0.062in}5 & \hspace{-0.005in}3 & \hspace{-0.11in}3 & 12 & \hspace{-0.065in}10 & \hspace{-0.113in}8(1) \\ Queen's & \hspace{0.062in}2 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & \hspace{0.062in}1 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & \hspace{0.062in}3 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & \hspace{0.062in}5 & \hspace{-0.005in}2 & \hspace{-0.11in}0 & 11 & \hspace{-0.005in}5 & \hspace{-0.113in}3 \\ York & \hspace{0.062in}2 & \hspace{-0.005in}0(1) & \hspace{-0.11in}0 & \hspace{0.062in}3 & \hspace{-0.005in}0 & \hspace{-0.11in}0 & \hspace{0.062in}1 & \hspace{-0.005in}0 & \hspace{-0.11in}0 & \hspace{0.062in}5 & \hspace{-0.005in}2 & \hspace{-0.11in}0 & 11 & \hspace{-0.005in}2(1) & \hspace{-0.113in}0 \\ Calgary & \hspace{0.062in}2 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & \hspace{0.062in}2 & \hspace{-0.005in}1 & \hspace{-0.11in}1 & \hspace{0.062in}1 & \hspace{-0.005in}1 & \hspace{-0.11in}0(1) & \hspace{0.062in}5 & \hspace{-0.005in}3 & \hspace{-0.11in}1 & 10 & \hspace{-0.005in}6 & \hspace{-0.113in}3(1) \\\hline Total & \hspace{-0.060in}216 & \hspace{-0.135in}131(2) & \hspace{-0.23in}104(21) & \hspace{-0.065in}184 & \hspace{-0.135in}116(1) & \hspace{-0.165in}84(14) & \hspace{-0.060in}225 & \hspace{-0.135in}157(2) & \hspace{-0.165in}86(22) & \hspace{-0.060in}272 & \hspace{-0.132in}179(3) & \hspace{-0.165in}48(14) & \hspace{-0.065in}897 & \hspace{-0.132in}583(8) & \hspace{-0.233in}322(71) \\\hline \end{tabular} \end{center} \end{table*} \normalsize The next, and most arduous, step in ultimately determining a given program's success rate, was determining the whereabouts and/or astronomy ``status'' of the 897 PhD recipients listed in column 14 of Table 1. The methodology employed was typically as follows: (i) ADS, the Science Citation Index (SCI), or the Institute for Scientific Information Citation Database (ISI)\footnote{The SCI and ISI are clearly superior for tracking down astronomy graduates whose publication habits, for one reason or another, avoid the traditional journals monitored by the ADS.} were searched for the most recent publication(s) - no preference was made as to an author's position in the author list; (ii) institutional affiliation and email address were noted; (iii) confirmation of institutional affiliation was ensured in all cases, either by relevant web page or email contact; (iv) institutional status (i.e., soft-money, fixed-term contract, tenured faculty, open-ended civil appointment, etc.) was confirmed in \it all \rm cases, again, either by web page or email contact. In total, $\sim 1/3$ of our total sample were contacted via email, the majority of whom provided clarification where necessary. No subjective classification based upon publication frequency was incorporated into the analysis; provided a present-day (late-1998) institutional affiliation could be confirmed, a graduate included in a given $n_{\rm g}$ entry in Table 1 would then also be counted in the corresponding $n_{\rm astr}$ entry. Occasionally, an unequivocal conclusion regarding a given graduate's status in the community was impossible; those falling into this category\footnote{Usually those who had either published a paper during 1997-1998, but for whom no confirmation of current institutional affiliation could be made, or those few who did not respond to the email request for institutional and/or job status clarification.} are included by their relevant entry of Table 1 in parantheses. Column 15 of Table 1 lists the number of 1975-1994 graduates still involved in astronomical research ($n_{\rm astr}$), for each of the institutions, while column 16 provides the number which are currently (as of late-1998) in identifiably permanent astronomy research positions. The latter are overwhelmingly based at university or national facilities, but occasionally may include industry positions if they are clearly related to astronomy research\footnote{Such positions are generally astronomical software or instrumentation development-related, the research for which is published, and not restricted to ``in-house'' documents.}. Again, five-year sub-totals are also provided under the relevant columns. \section{Discussion} \label{discussion} Before discussing any putative ranking based upon the production of long-term career astronomers, several general trends from Table 1 should be noted. First, the five-year sub-samples show that the number of 1985-1994 graduates (columns 8 and 11) is $\sim 25$\% greater than the number of 1975-1984 (columns 2 and 5) graduates, growth which, in North America, is reflected primarily in what were once the smaller North American programs (e.g., Hawaii, NMSU, Montreal, and Quuen's), a trend already noted by Domen \& Thronson (1988). \small \begin{table*} \caption{Fraction of PhD graduates still involved in astronomical research (f$_{\rm astr}$), as of late-1998, along with the fraction of graduates categorised as permanent or tenure-track (f$_{\rm perm}$).} \begin{center} \begin{tabular}{lp{0.45in}p{0.52in}p{0.45in}p{0.52in}p{0.45in}p{0.52in}p{0.45in}p{0.52in}p{0.45in}p{0.32in}}\hline Institution & \multicolumn{2}{c}{\hspace{-0.2in}1975-1979} & \multicolumn{2}{c}{\hspace{-0.2in}1980-1984} & \multicolumn{2}{c}{\hspace{-0.2in}1985-1989} & \multicolumn{2}{c}{\hspace{-0.2in}1990-1994} & \multicolumn{2}{c}{\hspace{-0.1in}Total} \\ & f$_{\rm astr}$ & \hspace{-0.05in}f$_{\rm perm}$ & f$_{\rm astr}$ & \hspace{-0.05in}f$_{\rm perm}$ & f$_{\rm astr}$ & \hspace{-0.05in}f$_{\rm perm}$ & f$_{\rm astr}$ & \hspace{-0.05in}f$_{\rm perm}$ & f$_{\rm astr}$ & \hspace{-0.05in}f$_{\rm perm}$ \\\hline Victoria & \hspace{-0.08in}1.00 & \hspace{-0.18in}1.00 & \hspace{-0.08in}1.00 & \hspace{-0.18in}1.00 & \hspace{-0.08in}1.00 & \hspace{-0.1in}.00-.50 & .60 & \hspace{-0.1in}.60 & \underline{.83} & \hspace{-0.1in}\underline{.67}-.75 \\ Montreal & .00 & \hspace{-0.1in}.00 & .00 & \hspace{-0.1in}.00 & \hspace{-0.08in}1.00 & \hspace{-0.1in}.00-.50 & .82 & \hspace{-0.1in}.09-.27 & \underline{.79} & \hspace{-0.1in}\underline{.07}-.29 \\ Harvard & .73 & \hspace{-0.1in}.62-.69 & .85 & \hspace{-0.1in}.65-.75 & .71 & \hspace{-0.1in}.42-.50 & .81 & \hspace{-0.1in}.24-.33 & \underline{.77} & \hspace{-0.1in}\underline{.48}-.57 \\ Hawaii & .83 & \hspace{-0.1in}.67-.83 & .67 & \hspace{-0.1in}.33 & \hspace{-0.08in}1.00 & \hspace{-0.1in}.67-1.0 & .71 & \hspace{-0.1in}.07 & \underline{.77} & \hspace{-0.1in}\underline{.31}-.38 \\ Princeton & .72 & \hspace{-0.1in}.56-.64 & .65 & \hspace{-0.1in}.35-.50 & .81 & \hspace{-0.1in}.56-.59 & .72 & \hspace{-0.1in}.34-.41 & \underline{.73} & \hspace{-0.1in}\underline{.46}-.53 \\ Maryland & .54 & \hspace{-0.1in}.35-.50 & .69 & \hspace{-0.1in}.31-.56 & .77 & \hspace{-0.1in}.23-.50 & .90 & \hspace{-0.1in}.25-.30 & \underline{.71} & \hspace{-0.1in}\underline{.29}-.46 \\ Caltech & .76 & \hspace{-0.1in}.55-.74 & .67 & \hspace{-0.1in}.55-.61 & .73 & \hspace{-0.1in}.42-.46 & .57 & \hspace{-0.1in}.22-.24 & \underline{.68} & \hspace{-0.1in}\underline{.43}-.52 \\ Colorado & .42 & \hspace{-0.1in}.42 & .80 & \hspace{-0.1in}.60-.70 & .67-.75 & \hspace{-0.1in}.33-.46 & .70-.80 & \hspace{-0.1in}.15-.20 & \underline{.65}-.71 & \hspace{-0.1in}\underline{.33}-.41 \\ Wyoming & .00 & \hspace{-0.1in}.00 & .60 & \hspace{-0.1in}.40-.60 & \hspace{-0.08in}1.00 & \hspace{-0.1in}.00 & .71 & \hspace{-0.1in}.14-.29 & \underline{.64} & \hspace{-0.1in}\underline{.21}-.36 \\ UBC & .63 & \hspace{-0.1in}.50-.63 & .33 & \hspace{-0.1in}.33 & \hspace{-0.08in}1.00 & \hspace{-0.1in}.33 & .57 & \hspace{-0.1in}.14 & \underline{.62} & \hspace{-0.1in}\underline{.33}-.38 \\ MSSSO & .67 & \hspace{-0.1in}.56 & .61 & \hspace{-0.1in}.33 & .56 & \hspace{-0.1in}.38 & .60 & \hspace{-0.1in}.12 & \underline{.60} & \hspace{-0.1in}\underline{.29} \\ Calgary & .50 & \hspace{-0.1in}.50 & .50 & \hspace{-0.1in}.50 & \hspace{-0.08in}1.00 & \hspace{-0.1in}.00-1.0 & .60 & \hspace{-0.1in}.20 & \underline{.60} & \hspace{-0.1in}\underline{.30}-.40 \\ Leiden & .61 & \hspace{-0.1in}.50-.61 & .64 & \hspace{-0.1in}.59 & .54 & \hspace{-0.1in}.38-.50 & .58 & \hspace{-0.1in}.08-.17 & \underline{.59} & \hspace{-0.1in}\underline{.38}-.46 \\ Sydney & .33 & \hspace{-0.1in}.33 & .33 & \hspace{-0.1in}.33 & .86 & \hspace{-0.1in}.36 & .53-.59 & \hspace{-0.1in}.12 & \underline{.57}-.59 & \hspace{-0.1in}\underline{.26} \\ NMSU & .50-.67 & \hspace{-0.1in}.33-.67 & .00-.50 & \hspace{-0.1in}.00-.50 & .67 & \hspace{-0.1in}.33-.44 & .60 & \hspace{-0.1in}.00-.20 & \underline{.55}-.64 & \hspace{-0.1in}\underline{.23}-.45 \\ Toronto & .47 & \hspace{-0.1in}.40 & .44 & \hspace{-0.1in}.33 & .53 & \hspace{-0.1in}.47 & .55 & \hspace{-0.1in}.09-.18 & \underline{.49} & \hspace{-0.1in}\underline{.34}-.36 \\ Western & .29 & \hspace{-0.1in}.29 & .00 & \hspace{-0.1in}.00 & .75 & \hspace{-0.1in}.25-.50 & .75 & \hspace{-0.1in}.25 & \underline{.47} & \hspace{-0.1in}\underline{.24}-.29 \\ Queen's & .50 & \hspace{-0.1in}.50 & \hspace{-0.08in}1.00 & \hspace{-0.18in}1.00 & .33 & \hspace{-0.1in}.33 & .40 & \hspace{-0.1in}.00 & \underline{.45} & \hspace{-0.1in}\underline{.27} \\ York & .00-.50 & \hspace{-0.1in}.00 & .00 & \hspace{-0.1in}.00 & .00 & \hspace{-0.1in}.00 & .40 & \hspace{-0.1in}.00 & \underline{.18}-.27 & \hspace{-0.1in}\underline{.00} \\\hline Total & .61-.62 & \hspace{-0.1in}.48-.58 & .63-.64 & \hspace{-0.1in}.46-.53 & .70-.71 & \hspace{-0.1in}.38-.48 & .66-.67 & \hspace{-0.1in}.18-.23 & \underline{.65}-.66 & \hspace{-0.1in}\underline{.36}-.44 \\\hline \end{tabular} \end{center} \end{table*} \normalsize Second, and of particular interest (concern?) for Australian astronomy, is the fact that MSSSO and Sydney have increased their number of PhD graduates by $\sim 50$\% and $\sim 100$\%, respectively, from 1975-1984 to 1985-1994. These two institutions dominate astronomy PhD production in Australia, accounting for 50\% of the total (Table 3.5.2 of ``Australian Astronomy: Beyond 2000''\footnote{After removal of the Monash University entry, which reflects the pure mathematics department as a whole, and not the small subset of astronomers therein.}). Of concern should be the fact that this $\sim 70$\% increase in the number of PhDs granted has not been matched by a parallel increase in the number of postdoctoral and permanent faculty/research positions. For comparison, during the same time period, the number of American, Canadian, and Dutch PhDs increased by the more modest (but not insubstantial) $\sim 20$\%. Shear PhD production (column 14 of Table 1), while (perhaps) a useful measure at some level, should not be the only yardstick by which a given program's career astronomer-production ``efficiency'' is measured. More useful is normalizing the present-day number of active (both soft-money and permanent) research astronomers produced by each institution ($n_{\rm astr}$ and $n_{\rm perm}$ of Table 1), by the number of PhDs awarded ($n_{\rm g}$). In Table 2, this fraction of total astronomy PhD graduates still involved in astronomy research f$_{\rm astr}$, along with the fraction possessing an identifiably permanent or tenured position f$_{\rm perm}$, are listed for each of the 19 institutions in our study. Columns 10 and 11 give the numbers based upon the entire 1975-1994 baseline, while the five-year sub-sample fractions are likewise tabulated in the appropriate columns. The parenthesized uncertainties of Table 1 are reflected by hyphenated ranges in Table 2 (i.e., the underlined lower limits of columns 10 and 11 are certain), but some leeway (particularly as far as f$_{\rm perm}$ goes) remains, due to unresolved permanence-vs-non-permanence issues, for some graduates. Sorting of the institutions in Table 2 was done, in descending order, by present-day fraction of astronomers still active in the field (i.e., column 10). Many caveats must be borne in mind when interpreting Table 2. First, and perhaps foremost, we must caution the reader against over-interpreting the rankings of the six small Canadian universities (the final six entrants of Table 1). With $<1$ PhD granted per year, during this 20-year period, their rankings are subject to the whims of small number statistics. Second, no provision (of course) is made for those who willingly left the field to pursue non-astronomy career goals; indeed, considering the present-day astronomy job market, and the PhD overproduction rate (Thronson 1991), programs which responsibly inform graduate students about the realities of the market, and offer parallel non-astronomy skills training, could end up suffering in the rankings in Table 2.\footnote{On the other hand, so very few institutions, in our experience, present such options, that the cynic might say that such a putative effect is entirely non-existent!} Third, it became readily apparent that certain schools have significant foreign student enrollment; due to the nature of \it some \rm overseas fellowships, this occasionally allows young PhD recipients to return to their home country to an early faculty/permanent position. Again, some schools have their f$_{\rm perm}$ enhanced by this mechanism, but in general the effect is small. A (perhaps) more impartial measure of f$_{\rm perm}$ would be to compare each institute's graduate fraction who find permanent employment in that institute's country; we will return to this point at the end of this section. Caveats of a more scientific nature are equally important to recall. First, averaging over any given institute's output can be potentially misleading; each school has their sphere of expertise, and at some level, one must worry about comparing an f$_{\rm astr}$ from an instrumentation-dominated program, with an f$_{\rm astr}$ from a theoretical cosmology program. Just as important, within any given school, there will exist a hierarchy in the success rates of particular PhD supervisors in training long-term career astronomers. This averaging over areas of expertise (and non-expertise) and individual faculty supervisors will not do justice to that lone supervisor who continually trains successful graduates, but who toils in relative obscurity in an otherwise mediocre department. Prospective students, for example, should bear these latter points in mind when investigating graduate school options, and should not blindly adhere to rankings of the like presented in Table 2, the annual U.S. News Gradute School or Gourman Rankings. Having presented the numerous caveats, though, there is still much to be learned from Table 2. Perhaps the most (pleasantly) surprising result of our analysis is that, with very few exceptions, \it $\sim 55\rightarrow 75$\% of all the graduates, regardless of original institutional affiliation, are still involved in astronomical research. \rm In total, 583 of the 897 PhD graduates in our study, or $\sim 65$\%, were still involved in research, \it at some level\rm, in late-1998. Examining the five-year sub-samples, we see that there is little variation in f$_{\rm astr}$, at the present-day, regardless of graduation date. The percentages for 1975-1979, 1980-1984, 1985-1989, and 1990-1994, are 61\%, 63\%, 70\%, and 66\%, respectively; we initially assumed the constancy in these f$_{\rm astr}$ values suggested that the decision to pursue non-astronomical interests was made within a few years of PhD receipt, with only marginal migration from the field thereafter. While this is perhaps true in some cases, we are now of the opinion that the similarity is more a conspiracy in the temporal evolution of the ``drop-out rate'', in that research ``half-life'' for graduates from the 1970s was longer than those from the 1990s, the result of which is the f$_{\rm astr}\approx 0.65$, for each of the five-year sub-samples of Table 2. The first three of the five-year sub-samples of Table 2 also show similar fractions of astronomers now in permanent positions f$_{\rm perm}$, typically $\sim 40\rightarrow 50$\%; apparently, within ten years of graduation, an equilibrium has been reached in which the ratio of astronomy graduates with permanent positions to graduates on soft-money to graduates who have left research entirely\footnote{Modulo our definition of ``leaving research'' described in Section \ref{method}.} is approximately 45:20:35. The most recent sub-sample (1990-1994), as witnessed by the final entry to Table 2, has clearly not reached this equilibrium, which is not surprising, since many of the graduates in question would still be in the midst of their second postdoctoral position. In passing, we note that the Canadian numbers are slightly lower than the aforementioned ${\rm f}_{\rm astr}=0.65$; $\sim 54$\% of Canadian graduates are still involved in astronomical research ($\sim 30$\% in permanent positions). Canada's marginally poorer performance, in this regard, should not be overly surprising, when one takes into account the dismal funding situation faced by Canadian astronomers (van der Kruit 1994). Regarding the implications for Australian astronomy - first, both the overall fraction still in astronomy ($\sim 59$\%) and the fraction with permanent positions ($\sim 28$\%), are only marginally lower than those for the entire sample ($\sim 65$\% and $\sim 36$\%, respectively). What is perhaps of some concern is that the fraction of graduates who have eventually found permanent positions \it within \rm Australia, as derived from Table 3, is only $\sim 11$\%; $\sim 20$\% of the 1975-1994 Sydney graduates currently fall into this category, while the fraction for MSSSO graduates is $\sim 6$\%. While this 6\% rate is one of the lowest of the 19 institutions in this study (Table 3), MSSSO compensates by training a very high fraction of graduates who eventually fill permanent overseas positions ($\sim 24$\%). In comparison, $\sim 17$\% of Canadian university astronomy graduates have settled into permanent positions within Canada, $\sim 21$\% of Leiden graduates now have permanent positions within the Netherlands, and $\sim 34$\% of American institutional graduates now have permanent positions with the USA. \small \begin{table}[ht] \caption{Fraction of PhD graduates with permanent, or tenure-track, astronomy research positions within the same country as their PhD institution ($f_{\rm perm}^{\rm same}$) - derived from number of graduates in this category ($n_{\rm perm}^{\rm same}$) and the total number of graduates ($n_{\rm g}$).} \begin{center} \begin{tabular}{lrrr}\hline Institution & $n_{\rm g}$ & $n_{\rm perm}^{\rm same}$ & $f_{\rm perm}^{\rm same}$ \\\hline Victoria & 12 & 7$\;\;$ & .58$\;\;$ \\ Harvard & 91 & 35$\;\;$ & .38$\;\;$ \\ Princeton & 101 & 38$\;\;$ & .38$\;\;$ \\ Caltech & 132 & 49$\;\;$ & .37$\;\;$ \\ Hawaii & 26 & 8$\;\;$ & .31$\;\;$ \\ Colorado & 66 & 20$\;\;$ & .30$\;\;$ \\ Maryland & 84 & 22$\;\;$ & .26$\;\;$ \\ NMSU & 22 & 5$\;\;$ & .23$\;\;$ \\ Wyoming & 14 & 3$\;\;$ & .21$\;\;$ \\ Leiden & 90 & 19$\;\;$ & .21$\;\;$ \\ Calgary & 10 & 2$\;\;$ & .20$\;\;$ \\ Sydney & 46 & 9$\;\;$ & .20$\;\;$ \\ UBC & 21 & 4$\;\;$ & .19$\;\;$ \\ Queen's & 11 & 2$\;\;$ & .18$\;\;$ \\ Western & 17 & 3$\;\;$ & .18$\;\;$ \\ Toronto & 61 & 7$\;\;$ & .11$\;\;$ \\ Montreal & 14 & 1$\;\;$ & .07$\;\;$ \\ MSSSO & 68 & 4$\;\;$ & .06$\;\;$ \\ York & 11 & 0$\;\;$ & .00$\;\;$ \\\hline Total & 897 & 238$\;\;$ & .27$\;\;$ \\\hline \end{tabular} \end{center} \end{table} \normalsize Another interesting aspect of the results of Tables 1-3, for Australian astronomy, relates back to one of the conclusions of the ``Australian Astronomy: Beyond 2000'' document (Section 2.2). A claim was made therein that $\sim 21$\% of Australian PhD graduates obtain permanent astronomy positions within Australia.\footnote{An estimate was made that 112 out of 160 Australian astronomy positions are permanent; with an assumed lifetime per position of 35 years, a claim for 3.2 permanent positions per year is made. With (typically) 15 PhDs granted per year, this leads to their inferred claim that 21\% of Australian graduates obtain permanent positions \it within \rm Australia.} Our analysis of the MSSSO and Sydney graduates, which, recall, account for half of \it all \rm Australian astronomy PhDs, shows that this estimate was a factor of two too optimistic. In retrospect this should not be too surprising, since an underlying assumption of the ``Australian Astronomy: Beyond 2000'' claim was that all permanent astronomy positions in Australia are filled by Australian graduates; anecdotally, this is clearly a flawed assumption, and one which we have now confirmed quantitatively. How do our results compare with the earlier Domen \& Thronson (1988) and Trimble (1991) analyses? The former clearly demonstrated that the subset of PhD graduates from the top five programs in the U.S. News Gradute School Rankings (Caltech, Princeton, Harvard, Berkeley, and Chicago), in comparison with five more moderately-ranked programs (Arizona, Colorado, UCLA, Minnesota, and Virginia), were disproportionately more successful (by more than a factor of six) in obtaining permanent research positions at the 32 largest PhD degree-granting institutions. Our survey was not designed to substantiate or repudiate Domen \& Thronson, but even a cursory analysis of our dataset would tend to lend credence to their claim, as would anecdotal wisdom - this is a negative aspect of such analyses. On the other hand, we choose to focus on the positive - our results clearly indicate that $\sim 65$\%$\pm 10$\% of graduates of \it any \rm astronomy PhD program can expect to maintain a career in research astronomy - i.e., actual \it source \rm of PhD is of little importance.\footnote{It is interesting to note that of the seven American universities ranked below Caltech (the top-ranked program according to the U.S. News Graduate School Rankings), the graduates of four of these (including two from outside the top ten in the U.S. News rankings) programs actually ranked (marginally) higher when it came to maintaining a long-term astronomy research career.} While it \it is \rm true, as Domen \& Thronson found, that the graduates of the so-called prestigious schools may preferentially fill positions at these same prestigious school, this does \it not \rm preclude the opportunity of pursuing a research career, quite often, for example, at one of the \it many \rm non-degree granting colleges and universities. While such positions generally carry a heavier burden of teaching and administrative duties, reduced publishing frequency, and lead some researchers to feel somewhat isolated from the community, they do still allow the determined astronomer to pursue a research career, albeit (perhaps) at a reduced efficiency from those in the large, active, degree-granting programs. This result is perhaps not fully appreciated upon initial reading of Domen \& Thronson, but is one in which the prospective graduating student should take heart! Our conclusions are in mild disagreement with those of Trimble (1991), although an \it a posteriori \rm examination of her dataset shows that the disagreement is not as striking as first thought. Recall first though, that Trimble found that 18 years after PhD receipt, graduates of one ``prestigious, top four'' (called `P') university were $\sim 60\rightarrow 100$\% more likely to be active in astronomy research than the graduates of one ``non-prestigious, second ten'' school (called `NP'). Our analysis though shows that for the eight US programs in our study, even the extrema (i.e., Harvard and NMSU) only differ 40\%; for the five large US programs (i.e., $>2$ PhD recipients per year), Harvard and Colorado differ by only 15\% (recall Table 2). In an attempt to uncover the source of the above discrepancy, we returned to Trimble's (1991) original Kepler-Meier survival curve (her Figure 2) showing the fraction of graduates from the two institutes in her study still publishing f$_{\rm pub}$, as a function of time since PhD receipt t$_{\rm PhD}$. While it is true that her dataset shows that graduates of `P' are a factor of two more likely to be publishing at t$_{\rm PhD}=18$\,yr, than are graduates of `NP', we feel this is an overly pessimistic reading of the data. The optimistic interpretation is that at t$_{\rm PhD}=16$\,yr the ``advantage'' enjoyed by graduates of `P' over those of `NP' is only $\sim 30$\%, consistent with our conclusions. At t$_{\rm PhD}=17$\,yr, the `NP' curve shows a precipitous drop in f$_{\rm pub}$, the significance of which should be tempered with the realization that limited numbers in the yearly bins have now become important. While Trimble stressed this t$_{\rm PhD}\ge 17$\,yr discrepancy in her conclusions, we would counter by stressing the similarity of the samples for t$_{\rm PhD}<17$\,yr. We are in a position to perform a similar Kaplan-Meier survival analysis for our sample, but now covering the 1975-1994 baseline. To parallel Trimble's (1991) study, let us contrast the behavior of the traditionally top-ranked program (Caltech), with that of one chosen from outside the top ten (Maryland). The fraction still publishing f$_{\rm pub}$, as a function of time since PhD receipt t$_{\rm PhD}$, for both programs, is represented in Figure 1 by the filled circles. \begin{figure}[ht] \centering \vspace{3.3in} \special{psfile=gibson_fig1.eps hoffset=-5 voffset=-70 vscale=43 hscale=43} \caption{Kaplan-Meier survival curves illustrating the fraction of PhD astronomers f$_{\rm pub}$ still publishing papers, as a function of time since PhD receipt t$_{\rm PhD}$. After Trimble (1991), a ``top four'' (Caltech) and ``second ten'' (Maryland) program are highlighted. Even after 20 years, using the full samples of Table 1 (filled circles), there is only a 30\% difference between the programs; more importantly, though, adopting the full sample has a tendency to mask recent trends. Since 1981, the fraction of Maryland graduates still involved in research is indistinguishable from the Caltech fraction; indeed, during the 1990s, one might argue that the roles have become reversed. \label{fig:fig1}} \end{figure} The first thing to note from Figure 1 is that we see no precipitous drop in f$_{\rm pub}$ beyond t$_{\rm PhD}=16$\,yr; over the full baseline, the difference in f$_{\rm pub}$ between the two programs is typically no more than 30\%. More importantly, Kaplan-Meier curves can mask (not surprisingly) trends occurring on timescales less than the baseline. This is particularly evident when contrasting the behavior of f$_{\rm pub}$, in the Maryland sample, since 1980; as the crosses in the upper panel demonstrate, this behavior is now indistinguishable from the Caltech sample, in the lower panel. In fact, since 1990, one might argue that Maryland graduates have been more successful in remaining in astronomy, although the numbers per bin are starting to get small, so we hesitate to overly interpret this result. A minor point to note is that $\sim 5\rightarrow 10$\% of graduates immediately leave the research field, never publishing anything beyond their PhD dissertation.\footnote{Unlike in Trimble (1991), we have counted the graduate's PhD dissertation as a publication, which is why that f$_{\rm pub}\equiv 1.0$ at t$_{\rm PhD}\equiv 0$\,yr.} While we have only shown two representative programs here (Caltech and Maryland), our ultimate conclusions do not depend on this choice, as was already evident by the institute-insensitive behavior of f$_{\rm astr}$ seen in Table 2. \section{Summary} \label{summary} Upon carefully analyzing the present-day status of 897 astronomy PhD recipients from 19 different institutions, our main conclusions can be summarized thusly: \begin{enumerate} \item Institutional source of one's PhD plays little part in whether or not (statistically) one will successfully pursue a long-term career in astronomy - $\sim 55\rightarrow 75$\% of all graduates, regardless of PhD source, do so. \item Graduates of traditionally ``prestigious'' programs are almost certainly disproportionately successful in obtaining permanent positions at similarly ranked schools (echoing Domen \& Thronson 1988), but this has obviously not precluded graduates of ``non-prestigious'' programs from pursuing their research careers from outside the elite, ranked, universities. Perhaps their efficiency has been hindered, but they have not had to completely forgo research either. \item Our results complement those of Domen \& Thronson (1988), although we have chosen to stress the positive aspects of the equal opportunity for long-term research careers, regardless of PhD source. Domen \& Thronson's sample is biased in that it is restricted to researchers who settle at the largest, degree-granting institutes, while our unbiased sample includes researchers at institutes large and small, foreign and domestic, degree- and non-degree-granting. \item ``Success'' is a dynamical entity - the survival analysis of Figure 1 demonstrates the danger of blindly adopting a 20-year baseline; in particular, recent trends will be masked. \item Within a decade of graduation, $\sim 45$\% of graduates will have an identifiably permanent or tenured position, $\sim 20$\% remain in the soft-money/fixed-term contract category, and $\sim 35$\% have left the field. \item American graduates are far more likely to obtain permanent astronomy positions in the USA ($\sim 34$\%), than are Dutch ($\sim 21$\%), Canadian ($\sim 17$\%), or Australian ($\sim 11$\%) graduates, in their respective countries. \item During the decade 1985-1994, the number of astronomy PhDs granted by US, Canadian, and Dutch institutes increased by $\sim$20\% over the previous decade; in Australia, though, the increase was $\sim$70\%. \end{enumerate} Again, if nothing else, the one point we wish to leave with the reader, or prospective graduate student, is the optimistic primary conclusion above - i.e., regardless of the source of your PhD, one has just as good a chance to pursue a research career in astronomy (albeit perhaps only at the part-time level), as the graduate of any other program. It is true that you may not have (statistically) the same odds as graduates of the traditional ``prestigious'' schools in pursuing this career at a large, degree-granting, university, but the research opportunities within smaller universities and colleges (both degree and non-degree granting), overseas institutes, government labs, and industry, appear to compensate. \acknowledgements The assistance of Don Faulkner, Gay Kennedy, John O'Byrne, and Tim de Zeeuw, in compiling the MSSSO, Sydney, and Leiden statistics, is gratefully acknowledged. We especially thank Virginia Trimble, for her careful reading of the manuscript, and the subsequent lively exchange of extremely helpful email! Finally, this paper may not have seen the light of day, without the aid of Kevin Marvel at the AAS; special thanks go out to him. This research has made \it extensive \rm use of NASA's Astrophysics Data System Abstract Service, the Science Citation Index, and the Institute of Scientific Information's Citation Database.
\section{Introduction} The deflection of light due to the gravitational field of matter inhomogeneities is observable through the distortion of images of background galaxies. In the case of weak gravitational fields, i.e., in the absence of strong gravitational lensing effects like giant arcs, the images of a population of background sources with known intrinsic ellipticity distribution can be used to statistically investigate weak gravitational lensing effects by measuring a net ellipticity. Cosmic shear -- the line-of-sight integrated tidal gravitational field -- reflects the statistical properties of the density fluctuation field (Gunn 1967, Blandford \& Jaroszynski 1981). A quantitative description of this connection is given by the two-point correlation function of galaxy-image ellipticities, or by the rms-ellipticity within a (circular) aperture, which was investigated by several authors using the linear and nonlinear power spectrum of density fluctuations in different cosmologies (see the recent review by Mellier 1998, and references therein). Schneider et al. (1998; hereafter SvWJK) investigated cosmic shear using the $M_{\rm ap}$-statistics which is a spatially filtered version of the projected density field. They computed the dispersion of $M_{\rm ap}$ for the linear and nonlinear power spectrum of density fluctuations in different cosmologies for a broad range of filter scales and showed that this quantity is a sensitive and `local' measure of the power spectrum; in fact, as shown by Bartelmann \& Schneider (1999), the mean-squared value of $M_{\rm ap}$ on an angular scale $\theta$ is a very good approximation to the power spectrum $P_\kappa$ of the projected density field at a wavenumber $s\approx 4.25/\theta$. SvWJK calculated the skewness of $M_{\rm ap}$ in the frame of quasi-linear theory of structure growth and found that it is a sensitive indicator for the density parameter $\Omega_0$, independent of the normalization of the power spectrum (see also Bernardeau, van Waerbeke \& Mellier 1997; van Waerbeke, Bernardeau \& Mellier 1999). The skewness measures the non-Gaussianity of the projected density field and indicates, according to its sign, an asymmetric positive or negative tail of the probability distribution function (PDF). The skewness of $M_{\rm ap}$, which reflects the skewness of the three-dismensional density fluctuations, was found to be positive, indicating an extended tail of the PDF towards high positive values of $M_{\rm ap}$. This behaviour is expected: the initial density contrast $\delta$, which is assumed to be a Gaussian random field, becomes non-Gaussian during its evolution through gravitational collapse. That means the PDF of $\delta$ attains a cut-off near $\delta=-1$ and a broad positive tail, according to the occurrence of underdense and overdense regions. Since $M_{\rm ap}$ is a linear function of the projected density field, the PDF of $M_{\rm ap}$ is closely related to that of $\delta$. Therefore the PDF of $M_{\rm ap}$ reflects the nonlinear evolution of $\delta$. The projected density field $\kappa$ in general is defined as a projection of the three-dimensional density contrast $\delta$, weighted by a redshift-dependent factor which accounts for the lensing geometry and the redshift distribution of the sources. The highest peaks of $\kappa$ are expected to arise from physical objects with high three-dimensional density, i.e., collapsed haloes. In Kruse \& Schneider (1999; hereafter KS99) we have calculated the number density of haloes which yield a value of $M_{\rm ap}$ larger than a certain threshold. We found that for all cosmologies considered, the number density of haloes above a threshold corresponding to a signal-to-noise of 5 exceeds ten per square degree, for currently feasible deep optical imaging, so that a wide-field deep optical survey should detect hundreds of such peaks, which can then be used to define a mass-selected sample of dark matter haloes (Schneider 1996). In this paper we continue the investigation of statistical properties of the aperture mass by computing its cumulative probability distribution function (CPDF) for large positive values of $M_{\rm ap}$, assuming that all of these are caused by collapsed structures. Describing the density profile of the haloes by the universal mass profile found by Navarro, Frenk \& White (1996, 1997; hereafter NFW), we can assign an angular cross-section to each halo. If this cross-section is integrated over the abundance of haloes as obtained from Press-Schechter theory, one can determine the probability to measure a value of $M_{\rm ap}$ larger than some threshold, i.e., the CPDF for the tail of $M_{\rm ap}$. Similar to the number density calculated in KS99, the amplitude and shape of this tail reflects the abundance of dark-matter haloes, which can be used as a powerful cosmological probe. The rest of this paper is organized as follows: in Sect.\ 2 we briefly review the concept of the aperture mass, as applied to the NFW- profile. The cross-sectional area as a function of $M_{\rm ap}$, and the CPDF, are derived in Sect.\ 3, and results are given in Sect.\ 5. The number density of peaks in the two-dimensional distribution of $M_{\rm ap}$ will be strongly affected, though in a controllable way, by noise, mainly coming from the intrinsic ellipticity distribution of galaxies. We consider an alternative observable for the abundance of peaks in $M_{\rm ap}$ in Sect.\ 4. Finally, we summarize and discuss our results in Sect.\ 6. \def{\mathrm d}{{\mathrm d}} \section{Formalism} Following Schneider (1996), we define the spatially filtered mass inside a circular aperture of angular radius $\theta$ around the point $\mbox{\boldmath$\zeta$}$, \begin{equation} M_{\rm ap} (\mbox{\boldmath$\zeta$}):=\int {\rm d}^2 \vartheta \ \kappa(\mbox{\boldmath$\vartheta$}) \ U(\vert \mbox{\boldmath$\vartheta-\zeta$} \vert), \label{mapt} \end{equation} where the continuous weight function $U(\vartheta)$ vanishes for $\vartheta>\theta$. If $U(\vartheta)$ is a compensated filter function, \begin{equation} \int_0^{\theta} {\rm d} \vartheta \ \vartheta \ U(\vartheta)=0, \end{equation} one can express $M_{\rm ap}$ in terms of the tangential shear $\gamma_{\rm t}(\mbox{\boldmath$\xi$}; \mbox{\boldmath$\zeta$})$ at position $\mbox{\boldmath$\xi + \zeta$}$ relative to the point $\mbox{\boldmath$\zeta$}$ \begin{equation} M_{\rm ap} (\mbox{\boldmath$\zeta$})=\int {\rm d}^{2} \xi \ \gamma_{\rm t} (\mbox{\boldmath$\xi$}; \mbox{\boldmath$ \zeta$}) \ Q(\vert \mbox{\boldmath$\xi$} \vert), \label{mapshear} \end{equation} where \begin{equation} \gamma_{\rm t}(\mbox{\boldmath$\xi$}; \mbox{\boldmath$\zeta$}) = -{\rm Re}(\gamma(\mbox{\boldmath$\xi+\zeta$}) e^{-2i \phi}) \label{tanshear} \end{equation} is the tangential component of the shear at relative position $\mbox{\boldmath$\xi$}=(\xi \ {\rm cos} \ \phi, \xi \ {\rm sin} \ \phi)$. The function $Q$ is related to $U$ by \begin{equation} Q(\vartheta) = \frac{2}{\vartheta^2} \ \int_0^{\vartheta} {\rm d} \vartheta^{\prime} \ \vartheta^{\prime} \ U(\vartheta^ {\prime}) \ - U(\vartheta) . \end{equation} We use a filter function from the family given in SvWJK, specifically we choose that with $l=1$. Then writing $U(\vartheta)=u(\vartheta/ \theta)/\theta^2$, and $Q(\vartheta)= q(\vartheta/\theta)/\theta^2$, \begin{equation} u(x)=\frac{9}{\pi} (1-x^2) \left( \frac{1}{3}-x^2 \right), \end{equation} and \begin{equation} q(x)=\frac{6}{\pi} x^2(1-x^2), \end{equation} with $u(x)=0$ and $q(x)=0$ for $x>1$. We will describe the mass density of dark matter haloes with the universal density profile introduced by NFW, \begin{equation} \rho (r)= \frac{3 H_0^2}{8 \pi G} \ (1+z)^3 \ \frac{\Omega_{\rm d}}{\Omega (z)} \ \frac{\delta_c}{r/r_{\rm s} (1+r/r_{\rm s})^2}, \end{equation} with \begin{equation} \Omega(z)=\frac{\Omega_{\rm d}}{a+\Omega_{\rm d} (1-a)+\Omega_{\rm v} (a^3-a)}, \ a=\frac{1}{1+z}. \end{equation} $\Omega_{\rm d}$ and $\Omega_{\rm v}$ denote the present-day density parameters in dust and in vacuum energy, respectively. Haloes identified at redshift $z$ with mass $M$ are described by the characteristic density $\delta_{\rm c}$ and the scaling radius $r_{\rm s}=r_{200}/c$ where $c$ is the concentration parameter (which is a function of $\delta_{\rm c}$) and $r_{200}$ is the virial radius, defined such that a sphere with radius $r_{200}$ has a mean interior density of $200 \ \rho_{\rm crit}$ and contains the halo mass $M_{200}$. We compute the parameters which specify the density profile according to the description in NFW using the fitting formulae given there. The surface mass density of the NFW-profile is given by (see Bartelmann 1996) \begin{equation} \Sigma (\vartheta)= \frac{6 H_0^2}{8 \pi G} \ (1+z)^3 \ \frac{\Omega_{\rm d}}{\Omega (z)} \ r_{\rm s} \ \delta_{\rm c} \ f \left( \frac{\vartheta}{\theta_{\rm s}} \right), \end{equation} with \begin{eqnarray} f(x)&=&\frac{1}{x^2-1} \nonumber \\ &\times \ \biggl\{&^{\displaystyle{ 1-\frac{2}{\sqrt{1-x^2}} \ \mbox{arctanh} \ \sqrt{\frac{1-x}{1+x}}, \ \mbox{for} \ x < 1}} _{\displaystyle{1-\frac{2}{\sqrt{x^2-1}} \ \mbox{arctan} \ \sqrt{\frac{x-1}{1+x}}, \ \mbox{for} \ x > 1}}, \end{eqnarray} and $\theta_{\rm s}= r_{\rm s}/D_{\rm d}$. $D_{\rm d}$ is the angular diameter distance to the lens. Introducing the critical surface mass density \begin{equation} \Sigma_{\rm cr}= \frac{c^2}{4 \pi G} \ \frac{D_{\rm s}} {D_{\rm d} D_{\rm ds}}, \label{geom} \end{equation} with $D_{\rm s}$ and $D_{\rm ds}$ being the angular diameter distances to the source and that from the lens to the source, we can define the dimensionless surface density (convergence) which is a function of source redshift \begin{equation} \kappa (\vartheta,z_{\rm d},z_{\rm s}) = \frac{\Sigma (\vartheta)} {\Sigma_{\rm cr}} = \kappa_0 \ f \left( \frac{\vartheta}{\theta_{\rm s}} \right), \end{equation} with \begin{equation} \kappa_0=3 \ (1+z)^3 \ \frac{\Omega_{\rm d}}{\Omega(z)}\ r_{\rm s} \ \frac{H_0^2}{c^2} \ \delta_c \ \frac{D_{\rm d} D_{\rm ds}}{D_{\rm s}}. \end{equation} We assume a normalized source redshift distribution of the form \begin{equation} p_z (z)= \frac{\beta}{z_0^3 \ \Gamma \left(\frac{3}{\beta} \right)} \ z^2 \ \mbox{exp}(-[z/z_{0}]^{\beta}), \label{sources} \end{equation} (see Brainerd et al. 1996). The mean redshift of this distribution is proportional to $z_{0}$ and depends on the parameter $\beta$ which describes how quickly the distribution falls off towards higher redshifts. We will use the values $\beta=1.5$ and $z_0=1$. For these values the mean redshift $\langle z \rangle$ is given by $\langle z \rangle =1.505 \ z_{0}$. With the distribution (\ref{sources}) we define a source distance-averaged surface density \begin{equation} \kappa (\vartheta,z_{\rm d}) = \int {\rm d}z_{\rm s} \ p_z (z_{\rm s}) \ \kappa (\vartheta,z_{\rm d},z_{\rm s}) = \bar \kappa_0 \ f \left( \frac{\vartheta}{\theta_{\rm s}} \right), \label{ka} \end{equation} with $\bar \kappa_0 = \int {\rm d} z_{\rm s} \ p_z (z_{\rm s}) \ \kappa_0$. Inserting (\ref{ka}) in (\ref{mapt}) we get, after introducing polar coordinates for $\mbox{\boldmath$\vartheta$}$ and setting $\mbox{\boldmath$\zeta$}=(\zeta,0)$, which is the radial distance of the aperture centre from the halo center, \begin{eqnarray} &&M_{\rm ap} (\zeta)= \int_{{\rm max}(0,\zeta-\theta)} ^{\zeta+\theta} {\rm d} \vartheta \ \vartheta \ \kappa(\vartheta) \ \nonumber \\ && \times \int_{-\phi_{\rm m}}^{\phi_{\rm m}} {\rm d} \phi \ U\left(\sqrt{\vartheta^2+\zeta^2-2 \vartheta \zeta {\rm cos} (2 \phi)}\right), \label{map} \end{eqnarray} where $\phi_{\rm m}= {\rm min} \left( \pi, {\rm arccos} \left(\frac{\vartheta^2+\zeta^2-\theta^2}{2 \vartheta \zeta}\right) \right)$. For the filter function chosen above , the $\phi$-integration can be performed analytically. In Figure \ref{map_zeta} we plot the aperture mass of a halo with mass $M=10^{15} M_{\odot}/h$ at redshift $z_{\rm d}=0.3$ as a function of the radial distance $\zeta$, using a filter scale of $\theta=2'$, for five cosmological models. If we fix the parameters $(M,z_{\rm d},\theta)$, the aperture mass is a monotonically decreasing function of $\zeta$ between $\zeta=0$ and the first root of $M_{\rm ap}$. \begin{figure} \center{\includegraphics[ width=\columnwidth, draft=false, ]{map_zeta.ps}} \caption{The aperture mass, as defined in (\ref{map}), as a function of the radial distance $\zeta$ computed for five different cosmologies as indicated by the line types. The numbers in parentheses are the normalization $\sigma_8$ annd the shape parameter $\Gamma$. The halo mass is $10^{15} M_{\odot}/h$ and its redshift $z_{\rm d}=0.3$. The filter radius is $\theta=2$ arcmin. \label{map_zeta}} \end{figure} \section{The tail of $M_{\rm ap}$-statistics} In this section we calculate the CPDF $P(>M_{\rm ap},\theta)$, i.e., the probability to find an aperture mass larger than $M_{\rm ap}$ using a filter with radius $\theta$. We concentrate on values of $M_{\rm ap}$ which are much larger than the rms value of $M_{\rm ap}$, i.e., we consider only the far tail of the probability distribution. We may assume that such high values of $M_{\rm ap}$ are caused exclusively by collapsed haloes. Thus, from the properties of such haloes, together with their abundance, we can then determine the CPDF. Assuming a halo characterized by its mass $M$ and redshift $z_{\rm d}$, we can invert the function (\ref{map}) for a given value of $M_{\rm ap}$ and a fixed filter radius $\theta$ (see Fig. \ref{map_zeta}). As a result we get the separation $\zeta=\zeta(M_{\rm ap}, \theta, M, z_{\rm d})$. Owing to the monotonic behaviour of $M_{\rm ap}$ between $\zeta=0$ and its first root, separations smaller than the one obtained by inversion correspond to aperture masses larger than the threshold $M_{\rm ap}$. Therefore, $\zeta$ defines for each halo an angular cross section \begin{equation} \sigma(M_{\rm ap},\theta,M,z_{\rm d}) := \pi \ \zeta^2 (M_{\rm ap},\theta,M,z_{\rm d}), \label{cro} \end{equation} which represents the halo target area for detecting a weak lensing signal with an aperture mass larger than $M_{\rm ap}$. The cross section (\ref{cro}) is non-zero if the aperture mass measured in the halo centre is larger than the threshold $M_{\rm ap}$ (see Fig. \ref{map_zeta}). The CPDF is now obtained by summing up the cross sections of all haloes within a unit solid angle; this yields \begin{eqnarray} &&P(>M_{\rm ap},\theta) = \frac{c}{H_0} \ \int {\mathrm d} z_{\rm d} \ \frac{ (1+z_{\rm d})^2}{E(z_{\rm d})} \ D_{\rm d}^2 (z_{\rm d}) \ \nonumber \\ &&\times \int {\rm d} M \ N_{\rm halo}(M,z_{\rm d}) \ \sigma(M_{\rm ap}, \theta,M,z_{\rm d}), \label{prob2} \end{eqnarray} with \begin{equation} E(z_{\rm d})= \sqrt{\Omega_{\rm d} (1+z_{\rm d})^3+ (1-\Omega_{\rm d} -\Omega_{\rm v})(1+z_{\rm d})^2+\Omega_{\rm v}}. \end{equation} $N_{\rm halo}(M,z_{\rm d}) \ {\rm d} M$ is the comoving number density of haloes with mass within ${\rm d} M$ about $M$ at redshift $z_{\rm d}$. We assume that $N_{\rm halo}(M,z_{\rm d})$ can be obtained from Press and Schechter (1974) theory (see, e.g. Lacey $\&$ Cole 1993, 1994). The $M$-integral in (\ref{prob2}) extends from a lower threshold $M_{\rm t}$ to infinity, where $\sigma$ vanishes for $M \leq M_{\rm t}$ (see KS99). \section{Number density of haloes, selected by $M_{\rm ap}$ and size} \begin{figure} \center{\includegraphics[ width=\columnwidth, draft=false, ]{zeta_m.ps}} \caption{The cross section radius obtained from inverting (\ref{map}) as a function of the lens mass for the same cosmological models as in Figure \ref{map_zeta}. (\ref{map}) is inverted for the aperture mass $M_{\rm ap}=0.04$, a filter radius $\theta=2$ arcmin and a halo redshift $z_{\rm d}=0.3$. The cross section of haloes characterized by the fixed parameters written in the panel is zero for halo masses smaller than $M_{\rm t} \sim 10^{13.8} M_{\odot}/h$. \label{zeta_m}} \end{figure} In KS99, we have calculated the number density of haloes $N(>M_{\rm ap},\theta)$ with an aperture mass $> M_{\rm ap}$, using the same physical model as described above. Considering $M_{\rm ap} (\theta)$ as a two-dimensional map, $N(>M_{\rm ap},\theta)$ yields, at fixed filter scale $\theta$, the number density of peaks in this map with amplitude $> M_{\rm ap}$. Such peaks may be generated by noise where the major contribution comes from the intrinsic ellipticity distribution of galaxies. Intuitively, one might expect that high peaks are less effected by noise than smaller ones; this motivates us to consider the number density of haloes with aperture mass $> M_{\rm ap}$ and cross-sectional area $>\sigma=\pi \zeta_{\rm t}^2$, which is again obtained from summing over all haloes per unit solid angle, \begin{eqnarray} &&N(> M_{\rm ap}, > \zeta_{\rm t}, \theta)= \frac{c}{H_0} \ \int {\mathrm d} z_{\rm d} \ \frac{(1+z_{\rm d})^2}{E(z_{\rm d})} \ D_{\rm d}^2 (z_{\rm d}) \ \nonumber\\ &&\times \int {\mathrm d} M \ N_{\rm halo}(M,z_{\rm d}) \ {\rm H}[\zeta(M_{\rm ap},\theta,M,z_{\rm d})-\zeta_{\rm t}], \label{numbercro} \end{eqnarray} where ${\rm H}$ is the Heaviside step function. The integrand is non-zero only for $M>M_{\rm t}$, where $M_{\rm t}=M_{\rm t} (\zeta_{\rm t}, z_{\rm d} ,M_{\rm ap},\theta)$ is the mass obtained by inversion of the function shown in Fig. \ref{zeta_m}. For $\zeta_{\rm t}=0$, $N(> M_{\rm ap}, > \zeta_{\rm t}, \theta)= N(> M_{\rm ap},\theta)$. The value of $\zeta_{\rm t}$ for which $N(> M_{\rm ap}, > \zeta_{\rm t}, \theta)$ decreases to about $1/2$ of $N(> M_{\rm ap},\theta)$ yields the characteristic size of peaks corresponding to a given $M_{\rm ap}$; this value is expected to be $< \theta$. E.g., for the EdS(0.6,0.25) cosmology we obtain $N(> 0.04, > 0.45', 2')= 4.6$ which can be compared to $N(> 0.04, 2')= 9.4$ computed in KS99. For this example the characteristic size of a peak is roughly $1/4$ of the filter radius. \section{results} In this section we consider $P(>M_{\rm ap},\theta)$ for different cosmological models in the regime where the PDF of $M_{\rm ap}$ describes the non-linear evolution of the density field. Furthermore, we use the observable (\ref{numbercro}) to get additional constraints for the cosmological parameters. We perform our calculations for the same five cosmological models as shown in Figure \ref{map_zeta}. For three of them, the power spectrum is approximately cluster normalized, which corresponds to $\sigma_8 \approx 0.6$ for an Einstein-de Sitter universe (EdS, $\Omega_{\rm d} = 1$, $\Omega_{\rm v} = 0$) and $\sigma_8 = 1$ for both an open universe (OCDM, $\Omega_{\rm d} = 0.3$, $\Omega_{\rm v} = 0$) and a spatially flat universe with cosmological constant ($\Lambda$CDM, $\Omega_{\rm d} = 0.3$, $\Omega_{\rm v} = 0.7$). For these models we use the shape parameter $\Gamma = 0.25$ which yields the best fit to the observed two-point correlation function of galaxies (Efstathiou 1996). The remaining two EdS models have higher normalization ($\sigma_8 = 1$, approximately corresponding to COBE normalization) or a different shape parameter ($\Gamma = 0.5$). \begin{figure*} \center{\includegraphics[ draft=false, ]{multi_prob.ps}} \caption{The probability as defined in (\ref{prob2}) for the filter radii $\theta=2,4,6,10$ arcmin. In each panel we plot the same cosmological models as indicated in Figs. 1 and 2. The aperture mass range is defined by $[M_0,2 M_0]$, where $M_0=5 \cdot \sigma_{\rm c} (\theta)$ [see (\ref{sn}) and (\ref{var})]. \label{multi_prob}} \end{figure*} In the absence of lensing, the expectation value of $M_{\rm ap}$ vanishes, and its dispersion depends on the intrinsic ellipticity distibution, as calculated in Schneider (1996), \begin{equation} \sigma_{\rm c} (\theta) = 0.016 \left( \frac{n}{ 30 \ {\rm arcmin}^{-2}} \right)^{-1/2} \left( \frac{\sigma_{\epsilon}}{0.2} \right) \left( \frac{\theta}{1'} \right)^{-1}. \label{var} \end{equation} We take $n=30$ arcmin$^{-2}$ and $\sigma_{\epsilon}=0.2$ for the number density of background galaxies and their intrinsic ellipticity distribution, respectively, as representative values in the following. The signal-to-noise ratio of $M_{\rm ap}$ is then defined as \begin{equation} S_{\rm c} (\theta) = \frac{M_{\rm ap} (\theta)}{\sigma_{\rm c} (\theta)}. \label{sn} \end{equation} In Fig. \ref{multi_prob} we plot the probability (\ref{prob2}) for four different filter scales for the cosmologies introduced in the beginning of this section, for values of $M_{\rm ap}$ between $M_0$ and $2M_0$, where the value of $M_0$ was chosen to correspond to a signal-to-noise ratio of five, $M_0=5 \sigma_{\rm c} (\theta)$. According to this figure the probability $P(>M_{\rm ap},\theta)$ can be well approximated by an exponential, \begin{equation} P(>M_{\rm ap},\theta) = p_0 \ {\rm exp} \left[- \frac{(M_{\rm ap}- M_0)}{c} \right], \label{fit} \end{equation} where $p_0$ and $c$ are fit parameters. From (\ref{fit}) it is clear that $p_0$ is the probability to find a value of $M_{\rm ap}$ larger than $M_0$. We determined $p_0$ and $c$ simply by assuming that the logarithm of $P(>M_{\rm ap},\theta)$ in the interval $[M_0,2 M_0]$ follows a straight line. The maximal relative deviation between the probability (\ref{prob2}) and its approximation (\ref{fit}) in Fig. \ref{multi_prob} is about 3 percent. \begin{figure*} \center{\includegraphics[ draft=false, ]{multi_2_survey.ps}} \caption{The fit parameters of the exponential (\ref{fit}) denoted by crosses computed as described in the text together with their 1-$\sigma$ error ellipses which are defined by the dispersions (\ref{disp_c}) and (\ref{disp_p0}). For all panels the signal-to-noise ratio threshold $M_0/\sigma_{\rm c}$ is 5. Haloes with redshift in $z_{\rm d} \in [0,1]$ are considered. We use the source redshift distribution (\ref{sources}) with $\beta=1.5$ and $z_0=1$. The thin and the thick curves describe a 25 deg$^2$ and a 100 deg$^2$ survey, respectively. The cosmological models are indicated in the upper left panel. The different panels are obtained by varying the filter scale, $\theta=1,2,3,5$ arcmin. We do not plot the EdS(1.0,0.25) models because it is well separated from the other cosmologies ($p_0$ is about a factor of 5 larger compared to the probabilities of the remaining model). \label{multi_2_survey}} \end{figure*} \begin{figure*} \center{\includegraphics[ draft=false, ]{multi_mix_new.ps} \caption{The {top left} plot is the same as in Fig. \ref{multi_2_survey} for $L=10$ deg. In the {top right} plot all parameters are unchanged but only halo redshifts $z_{\rm d} \in [0.6,1]$ are considered. The {bottom left} plot is the same as the {top right} one but halo redshifts are from the interval $z_{\rm d} \in [0,0.4]$. In the {bottom right} plot all sources are located at redshift $z_{\rm s}=3$; the other parameters are the same as in the {top left} plot. As in Fig. \ref{multi_2_survey} we do not plot the high normalized EdS model which again is well separated from the other models. In the top right panel the value of $p_0$ for the EdS(0.6,0.25) model is so small that the normalization constant $f$ differs from unity appreciably; we have therefore not included this model in the figure. The cosmological models are indicated in the upper left panel}. \label{multi_mix_new}} \end{figure*} If we compute the probability according to the approximation (\ref{fit}), all information on cosmology is contained in the parameters $p_0$ and $c$. If we attempt to constrain the cosmological parameter set, we have to compute dispersions for the fit parameters. In the following we employ a maximum likelihood analysis to derive confidence levels in the parameter space $\{c,p_0\}$. In order to obtain a likelihood function we have to specify a probability distribution for finding a particular set of values $\{ M_{\rm ap}^{\rm i} \}$, ${\rm i} \in [1,N_{\rm f}]$ of $N_{\rm f}$ statistically independent aperture mass values. As an example, we shall assume that we have an image with side length $L$. SvWJK showed that about $N_{\rm f}= (L/(2 \theta))^2$ statistically independent fields can be placed on the image. This means that we consider two fields to be statistically independent if their angular separation is about two filter radii since then the correlation coefficient of the two fields is below 1 percent (see SvWJK). We suppose that this sample contains $N_>$ values of $M_{\rm ap}$ which are above the threshold $M_0$. This subsample consists of aperture mass values $M_{\rm ap}^{\rm j}>M_0$, ${\rm j} \in [1,N_>]$. Now we may ask for the probability to find $N_> \in [1,N_{\rm f}]$ values above the threshold if we have a sample of $N_{\rm f}$ fields. For each of the $N_{\rm f}$ aperture mass measurements the event $M_{\rm ap} > M_0$ is supposed to occur with probability $p_0$. Obviously this random process can be described by a binomial distribution, \begin{equation} P(N_>| N_{\rm f},p_0) = { N_{\rm f} \choose N_> } \ p_0^{N_>} \ (1-p_0)^{N_{\rm f}-N_>}. \label{bino} \end{equation} Of course, we can hope to estimate the two parameters $p_0$ and $c$ only for observations for which $N_> \gg 1$; in particular, if $N_> =0$, the parameter $c$ will be completely undetermined. Thus, we want to consider only realizations of (\ref{bino}) with $N_> \ge 1$, and therefore renormalize $P(N_>| N_{\rm f},p_0)$, \begin{equation} f \sum_{N_>=1}^{N_{\rm f}} \ P(N_>| N_{\rm f},p_0) = 1, \end{equation} where \begin{equation} f = \frac{1}{1-(1-p_0)^{N_{\rm f}}}. \label{f} \end{equation} In addition, we consider the probability distribution of $M_{\rm ap}$ itself for values above the threshold $M_0$. The normalized PDF of $M_{\rm ap} \ge M_0$ can be obtained from (\ref{fit}) by \begin{eqnarray} \hat p(M_{\rm ap})&&=\frac{1}{p_0} \ \left| \frac{{\rm d}}{{\rm d} M_{\rm ap}} P(>M_{\rm ap},\theta) \right| \nonumber \\ &&= \frac{1}{c} \ {\rm exp} \left[- \frac{(M_{\rm ap}- M_0)}{{\rm c}} \right]. \label{pdf} \end{eqnarray} Combining both probabilities, we obtain the likelihood function \begin{equation} {\rm L} (M_{\rm ap}^1, ..., M_{\rm ap}^{N_>}| p_0, c) = P(N_>| N_{\rm f},p_0) \ \prod_{{\rm j}=1}^{N_>} \ \hat p(M_{\rm ap}^{\rm j}), \label{likeli} \end{equation} which describes the probability to find $N_>$ aperture masses $>M_0$ where those above the threshold follow the distribution (\ref{pdf}). From (\ref{likeli}) we can derive maximum likelihood estimates for both fit parameters. Denoting these estimates by $(\hat c, \hat p_0)$ we obtain by differentiating (\ref{likeli}) w.r.t. $c$ and $p_0$ \begin{equation} \hat c = \frac{1}{N_>} \ \sum_{j=1}^{N_>} \ M_{\rm ap}^{\rm j} - M_0 \label{c} \end{equation} and \begin{equation} \hat p_0 = \frac{N_>}{N_{\rm f}}; \label{p} \end{equation} these results are of course not unexpected. In order to show that (\ref{c}) and (\ref{p}) are unbiased estimators we have to evaluate their ensemble average which can be performed by applying the operator \begin{eqnarray} {\rm P} (X) &=& f \sum_{N_>=1}^{N_{\rm f}} \ { N_{\rm f} \choose N_> } \ p_0^{N_>} \ (1-p_0)^{N_{\rm f}-N_>} \nonumber \\ &\times&\prod_{{\rm j}=1}^{N_>} \ \int_{M_0}^{\infty} {\rm d} M_{\rm ap}^{\rm j} \ \hat p(M_{\rm ap}^{\rm j}) \ \ ( X ). \label{op} \end{eqnarray} Averaging (\ref{c}) and (\ref{p}) with (\ref{op}) yields \begin{equation} \langle\hat c\rangle = {\rm P} (\hat c) = c \ {\rm and} \ \langle\hat p_0\rangle = {\rm P} (\hat p_0) = p_0 f. \label{mest} \end{equation} Because of the second of eqs.(\ref{mest}), $\hat p_0$ is an unbiased estimator for $p_0$ only if $f \approx 1$. In the following we will consider only values of $p_0$ and $N_{\rm f}$ which guarantee that $f \approx 1$ is satisfied, otherwise the statistics is not good enough to determine the parameters accurately anyway. The correlation between the two estimators can be calculated from \begin{equation} \langle\hat c \ \hat p_0\rangle = {\rm P} (\hat c \ \hat p_0) = c \ p_0 \ f. \label{corr} \end{equation} Since we have $f \approx 1$ these two estimators are not correlated. In the following we set $f=1$. Dispersions of the estimators are defined by \begin{equation} \sigma_{\hat p_0} = \sqrt{\langle \hat p_0^2 \rangle - \langle \hat p_0 \rangle^2} \end{equation} and \begin{equation} \sigma_{\hat c} = \sqrt{\langle \hat c^2\rangle - \langle\hat c\rangle^2}. \end{equation} The calculation of the dispersions involves the application of the operator (\ref{op}) to the squared estimators. After some algebra we obtain \begin{equation} \langle\hat c^2\rangle = c^2 \ \left(1 + \sum_{N_>=1}^{N_{\rm f}} \ P(N_{\rm f},p_0) \ N_>^{-1}\right) \end{equation} and \begin{equation} \langle \hat p_0^2 \rangle = \frac{p_0}{N_{\rm f}} (1 - p_0) + p_0^2, \end{equation} which together with (\ref{mest}) leads to \begin{equation} \sigma_{\hat c} =c \ \sqrt{\sum_{N_>=1}^{N_{\rm f}} \ P(N_>| N_{\rm f},p_0) \ N_>^{-1}}, \label{disp_c} \end{equation} and \begin{equation} \sigma_{\hat p_0}= \sqrt{\frac{p_0}{N_{\rm f}} (1 - p_0)}. \label{disp_p0} \end{equation} As expected, the rms values of the estimators decrease with increasing image size. This behaviour is obvious for (\ref{disp_p0}) and can be seen for (\ref{disp_c}) if we use a recurrence relation to compute the binomial distribution for a given $N_>$. The offset for the recurrence relation, $C(1) = N_{\rm f} p_0 (1-p_0)^{N_{\rm f}-1}$, rapidly decreases for large $N_{\rm f}$ and therefore the sum in (\ref{disp_c}). Note that the rms values of the estimators depend only on their means and $N_{\rm f}$ which is determined by the image size for a given filter scale. Given that the mean values $(c, p_0)$ depend on the filter radius $\theta$ and the threshold $M_0$ we can vary the parameter triple $(\theta,M_0,N_{\rm f})$ to obtain a significant difference between the various cosmological parameters. Furthermore, redshift information coming from possible redshift measurements of the haloes and the sources can be used to maximize the difference between various cosmologies. In Fig. \ref{multi_2_survey} we plot the mean values of the fit parameters (denoted by crosses) for the filter radii $\theta=1,2,3,5$ arcmin for four cosmological models. The threshold $M_0$ is defined by a signal-to-noise ratio of 5. For each cosmology and filter scale we use a 25 deg$^2$ (thin lines) and 100 deg$^2$ (thick lines) survey to compute the dispersions of the estimators (\ref{p}) and (\ref{c}) which define the 1-$\sigma$ error ellipses in Fig. \ref{multi_2_survey}. As expected, if we double the side length of the image the dispersions of $c$ and $p_0$ become smaller by about a factor of two. The COBE normalized EdS model is well separated for all filter scales. Since the probability $p_0$ in this model is about a factor of 5 larger than that of the remaining cosmologies we do not plot this model. The possibility of distinguishing cosmologies strongly depends on the filter scale. For an aperture with 1 arcmin filter radius we can clearly distinguish the EdS(0.6,0.25) and the low density model without cosmological constant from the other cosmologies if we use the survey with the larger area. If we enlarge the filter radius, the differences between the two low-density models and the EdS model with large shape parameter become smaller whereas the EdS(0.6,0.25) model remains distinguishable from all the other cosmologies. In Fig. \ref{multi_mix_new} we demonstrate the effects on the dispersions if we change the halo redshift integration range and the source redshift $z_{\rm s}$, where we assume all sources to be located at the same redshift. For reference, we choose the upper left plot from Fig. \ref{multi_2_survey} with $L=10$ deg. If we shrink the halo redshift interval to $z_{\rm d}=[0.6,1]$ we obtain the upper right plot. Because of the stronger evolution of the halo number density for high redshifts in the other cosmologies, the EdS(0.6,0.5) model can now be distinguished from the low-density models. The EdS(0.6,0.25) model has a value $p_0$ which is so small as to lead to a significant deviation of $f$ -- as defined in (\ref{f}) -- from unity, so that the expressions (\ref{disp_c}) and (\ref{disp_p0}) are no longer valid, and therefore we have not plotted this model. In the bottom left plot, compared to the upper right one, we only consider haloes having redshifts in the interval $z_{\rm d}=[0, 0.4]$. In this redshift range all cosmologies considered are distinguishable. The reason for the low-density models now being separated is the increasing difference between the rich cluster mass functions of both models if we choose smaller redshifts. In the remaining plot, in comparison with the top left one, all sources are assumed to have the same redshift $z_{\rm s}=3$ which is about twice the mean source redshift used in the other panels. Because of the improved efficiency of the weak lensing signal, the probability is increased by about a factor of 4. For this very deep survey all cosmologies are clearly separated. From Figs. \ref{multi_2_survey} and \ref{multi_mix_new} it becomes clear that we need large deep surveys in order to be able to distinguish clearly between the cosmological models considered, using only the statistics applied here. If we want to compute $P(>M_{\rm ap},\theta)$ as expected from real observations, we have to consider possible sources of error which may change the theoretical value of $M_{\rm ap}$, because the cross sections are given via aperture mass measurements (see Fig. \ref{map_zeta}). To account for the intrinsic ellipticity distribution of galaxies, we can proceed in the same way as for the observable $N(>M_{\rm ap}, \theta)$ in KS99. Essentially this means that we convolve $P(>M_{\rm ap},\theta)$ with a Gaussian defined by the dispersion of the intrinsic galaxy ellipticity distribution of the sources. As explained in SvWJK, this dispersion is expected to dominate the noise in a measurement of $M_{\rm ap}$. The probability obtained from the convolution can be directly compared with observations. We have checked that, as expected from the results of KS99, the convolution only slightly enhances the theoretical values and does not change the shape of $P(>M_{\rm ap},\theta)$ appreciably in the range of $M_{\rm ap}$ considered here. Therefore the fit formula we obtained remains valid, and the maximum likelihood method can be applied to derive dispersions for the fit parameters. In Table \ref{table} we show the number density of haloes with aperture mass larger than $M_{\rm ap}=0.04$ and cross-section radii exceeding the threshold $\zeta_{\rm t}=0.8$ arcmin for the filter radius $\theta=2$ arcmin. According to the fact that the rich cluster mass function shows most clearly the cosmology dependence of the non-linear evolution of the density field and since cross-section radii above 0.8 arcmin correspond to the most massive non-linear objects (see Fig. \ref{zeta_m}), $N(>M_{\rm ap}, > 0.8', \theta)$ better distinguishes cosmologies than $N(>M_{\rm ap}, \theta)$ for the same $M_{\rm ap}$ and $\theta$ (see KS99). A drawback when using the observable (\ref{numbercro}) is the larger image size required for the detection of significant differences between various cosmologies. From the numbers in Table \ref{table} we infer a survey area of 25 deg$^2$ which is needed to distinguish the cosmologies considered here significantly, i.e., with no overlapping Poissonian error bars. \begin{table} \caption{ The number of haloes per square degree with aperture mass larger than $M_{\rm ap}=0.04$ and cross-section radius greater than $\zeta_{\rm t}=0.8$ arcmin, as defined in (\ref{numbercro}), computed for five cosmological models. The filter radius is $\theta=2$ arcmin. The aperture mass and the dispersion determined by the filter radius correspond to a signal-to-noise ratio of 5. } \label{table} \begin{tabular}{lccccc} \noalign{\smallskip}\hline\noalign{\smallskip} $\Omega_0$ & $\Lambda$ & $\Gamma$ & $\sigma_8$ & $N(> 0.8',> 0.04,2')$ \\ \noalign{\smallskip}\hline\noalign{\smallskip} 1.0 & 0 & 0.25 & 0.6 & 0.47 \\ 1.0 & 0 & 0.25 & 1.0 & 11.9 \\ 1.0 & 0 & 0.5 & 0.6 & 1.16 \\ 0.3 & 0 & 0.25 & 1.0 & 2.22 \\ 0.3 & 0.7 & 0.25 & 1.0 & 1.67 \\ \noalign{\smallskip}\hline\noalign{\smallskip} \end{tabular} \end{table} \section{Discussion and Conclusions} In this paper we computed the highly non-Gaussian tail of the probability distribution of the aperture mass $M_{\rm ap}$ resulting from lensing by the large-scale structure. The CPDF is obtained by summing up the cross sections of dark matter haloes with assumed spherical mass density following a universal NFW density profile. We used Press-Schechter theory to compute the number density of massive haloes which cause the extended tail in the PDF of $M_{\rm ap}$. The number density of haloes with large cross-sectional radii, or in other words, with large masses, is a sensitive measure for constraining the cosmological parameter set (see KS99, and references therein). We showed that the number density in a mass-selected sample of haloes with cross sectional radii above 0.8 arcmin, which corresponds to halo masses $\sim 10^{14} M_{\odot}/h$, is measureable in all cosmologies considered. Especially, from a deep, high-quality imaging survey of 25 deg$^2$, some of the currently most popular cosmologies can be distinguished by the varying number of haloes with the selected cross section threshold. The expected number of massive haloes varies from $\sim 12$ for the EdS(0.6,0.25) to $\sim 300$ for the high normalised EdS model if we use the above-mentioned survey. The main result of this work is an analytical formula describing the PDF of the aperture mass, which turns out to be closely approximated by an exponential for values of $M_{\rm ap}$ well above its rms. This fact allows a simple maximum-likelihood analysis for distinguishing various cosmological models, using the non-Gaussian tail of $M_{\rm ap}$ only, as demonstrated in Sect.\ 5. Modifications of the present investigation can well be imagined and may turn out to yield fruitful results in future. One is the use of (photometric) redshift estimates which allow a more precise measurement of the shear, owing to the redshift-dependent geometrical factors entering the projected surface mass density. If the haloes found by our method turn out to be associated with galaxy concentrations, their redshift can be estimated, and the halo abundance as a function of $M_{\rm ap}$ and redshift can be obtained, allowing to greatly refine the statistics considered here and in KS99. Our approach to attempt measuring cosmological parameters is complementary to investigations employing two-point statistical measures, such as the rms shear in (circular) apertures or the shear two-point correlation function (e.g., Blandford et al. 1991; Kaiser 1992, 1998; Villumsen 1996; Jain \& Seljak 1997; Bartelmann \& Schneider 1999), and those which consider the skewness of cosmic shear statistics as a powerful handle on $\Omega_0$ (e.g., Bernardeau et al. 1997; van Waerbeke et al. 1999; Jain et al. 1999). In contrast to these lower-order statistical measures, the highly non-Gaussian features of cosmic shear are expected to be less affected by noise, whose main contribution is the intrinsic ellipticity dispersion of source galaxies. We note that the cosmology dependence of our results are mainly (but not exclusively) through the abundance of dark matter haloes as a function of mass and redshift. Thus, our statistics provides a fairly direct measure of this cluster abundance. In a future work, we shall attempt to relate $M_{\rm ap}$ to the true three-dimensional mass of haloes as determined in numerical $N$-body simulations, and thus relate the aperture mass statistics directly to the mass function of haloes. It should be stressed that, whereas the aperture mass is most likely not the optimal statistics to measure the cosmic shear power spectrum (see Kaiser 1998, and Seljak 1998, for different approaches), it is a particularly convenient measure for highly non-Gaussian, spatially localized features which can be obtained locally and directly from the observed image ellipticities and whose noise properties can be straightforwardly investigated. In particular, cosmic shear measurement using $M_{\rm ap}$, and the search for mass-selected haloes (Schneider 1996; KS99) are just two aspects of the same underlying physics and statistics. The validity of the various approximations which enter KS99 and the current study needs to be investigated in more detail. In a forthcoming paper (Reblinsky et al., in preparation) we will apply the aperture mass statistics to the same numerical simulations used in Jain et al.\ (1999), which combine very large N-body simulations with ray-tracing methods. Such numerical simulations are indispensible not only to assess the accuracy of the analytical approximation, but also to study statistical estimators which cannot be calculated analytically. Given the highly non-Gaussian nature of the projected density field resulting from the evolved LSS density distribution, it is by no means clear how to optimally and robustly distinguish between different cosmological models. The use of the far tail of the $M_{\rm ap}$-statistics should be viewed as one of several useful tools. \section*{Acknowledgments} This work was supported by the ``Sonderforschungsbereich 375-95 f\"ur Astro-Teilchenphysik" der Deutschen Forschungsgemeinschaft. We want to thank B. Geiger and L. van Waerbeke for many interesting discussions, and M. Bartelmann for a careful reading of the manuscript.
\section{Introduction } \label{sec-introd} \andy{intro} Highly non-classical, Schr\"odinger-cat-like neutron states can be produced by coherently superposing different spin states in an interferometer and with neutron spin echo \cite{RS,RSP}. We analyze here an interesting recent experiment \cite{BRSW} in which a polarized neutron crosses a magnetic field that is orthogonal to its spin, producing Schr\"odinger-cat-like states. Our main purpose is to investigate the decoherence effects that arise when the fluctuations of the magnetic field are considered. \section{Squeezing and squashing} \label{sec-sqesqa} \andy{sqesqa} Let us start by looking at the coherence properties of a neutron wave packet and concentrate our attention on the losses of coherence provoked by a fluctuating magnetic field. To this end, we introduce the Wigner quasidistribution function \andy{wigdef} \begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:wigdef} W(x,k) = \frac{1}{2\pi}\int d\xi\; e^{-ik\xi} \psi \left(x+\frac{\xi}{2} \right) \psi^*\left(x-\frac{\xi}{2} \right) , \eeq where $x$ is position, $p=\hbar k$ momentum and $\psi$ the wave function of the neutron in the apparatus. The Wigner function is normalized to one and its marginals represent the position and momentum probability distributions \andy{margxp} \begin{equation}}\newcommand{\eeq}{\end{equation} \int dx\; dk\; W(x,k) =1; \qquad P(x) = \int dk\; W(x,k), \quad P(k) = \int dx\; W(x,k). \label{eq:margxp} \eeq We shall work in one dimension. We assume that the neutron wave function is well approximated by a Gaussian \andy{gauss,gaussinv} \begin{eqnarray}}\newcommand{\earr}{\end{eqnarray} \psi(x) &=& \frac{1}{(2\pi\delta^2)^{1/4}} \exp \left[-\frac{(x-x_0)^2}{4\delta^2} + i k_0 x\right] , \label{eq:gauss} \\ \phi(k) &=& \frac{1}{(2\pi\delta_k^2)^{1/4}} \exp \left[-\frac{(k-k_0)^2}{4\delta_k^2} - i(k-k_0)x_0\right] \nonumber \\ &=& \left(\frac{2\delta^2}{\pi} \right)^{1/4} \exp \left[-\delta^2(k-k_0)^2 - i(k-k_0)x_0\right], \label{eq:gaussinv} \earr where $\delta$ is the spatial spread of the wave packet, $\delta_k \delta= 1/2$, $x_0$ is the initial average position of the neutron and $p_0=\hbar k_0$ its average momentum. The two functions above are related by a Fourier transformation and are both normalized to one. Normalization will play an important role in our analysis and will never be neglected. The Wigner function for the state (\ref{eq:gauss})-(\ref{eq:gaussinv}) is readily calculated \andy{wiggauss} \begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:wiggauss} W(x,k) = \frac{1}{\pi} \exp \left[-\frac{(x-x_0)^2}{2\delta^2}\right] \exp \left[-2\delta^2(k-k_0)^2\right] \eeq and turns out to be a positive function. In the language of quantum optics \cite{QOpt}, we shall say that the neutron is prepared in a coherent state if $\delta = \delta_k= 1/\sqrt{2}$ and in a squeezed state if $\delta \neq \delta_k$. An illustrative example is given in Figure 1. \begin{figure} \centerline{\epsfig{file=squasque.eps, width=\textwidth}} \caption{Fig. 1. Above, Wigner function (\ref{eq:wiggauss}), for $x_0=0, k_0=1.7 \cdot 10^{10}$m$^{-1}$; $x$ is in units $10^{-10}$m and $p=k$ in units $10^{10}$m$^{-1}$. From left to right: $\delta= 1/\sqrt{2}$ (coherent state), $\delta=1$ (squeezed state) and $\delta = \sqrt{2}$ (a more squeezed state); the uncertainty principle always reads $\delta_k\delta=1/2$ (minimum uncertainty states). Below, Wigner function in (\ref{eq:wigm}) for the same values of $x_0,k_0$ and $\Delta_0=0$. From left to right: $\sigma=0,1/\sqrt{2},\sqrt{3/2}$; the uncertainty principle yields $\delta_k\delta'=1/2,1/\sqrt{2},1$} (squashing). \end{figure} Consider now a polarized neutron that crosses a constant magnetic field, parallel to its spin, of intensity $B$ and contained in a region of length $L$. Since the total energy is conserved, the kinetic energy of the neutron in the field changes by $\Delta E = \mu B >0$, where $-\mu$ is the neutron magnetic moment. This implies a change in average momentum $\Delta k= m\mu B /\hbar^2 k_0$ and an additional shift of the neutron phase proportional to $\Delta \equiv L\Delta k/k_0$. The resulting effect on the Wigner function is $W(x,k)\to W(x-\Delta,k)$. Assume now that the intensity of the $B$-field fluctuates around its average $B_0$ according to a Gaussian law. This fluctuation is reflected in a fluctuation of the quantity $\Delta$ according to the distribution law \andy{probdel} \begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:probdel} w(\Delta) = \frac{1}{\sqrt{2\pi\sigma^2}} \exp \left[-\frac{(\Delta-\Delta_0)^2}{2\sigma^2}\right], \eeq where $\sigma$ is the standard deviation. The ratio $\sigma/\Delta_0$ is simply equal to the ratio $\delta B/B_0$, $\delta B$ being the standard deviation of the fluctuating $B$-field. The average Wigner function, when the neutron has crossed the whole $B$ region of lenght $L$, represents a ``squashed" state, that has partially lost its quantum coherence: \andy{wigm} \begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:wigm} W_{\rm m}(x,k)= \int d\Delta\; w(\Delta)\; W(x-\Delta,k). \eeq This function is represented in Figure~1 for $\Delta_0=0$ (vanishing average magnetic field) and increasing values of $\sigma$. The above Wigner function can be calculated explicitly, but its expression is a bit cumbersome; however, its marginals (\ref{eq:margxp}) are simple: \andy{marg1,marg2} \begin{eqnarray}}\newcommand{\earr}{\end{eqnarray} P(x) &=& \frac{1}{\sqrt{2\pi(\delta^2+\sigma^2)}} \exp \left[-\frac{(x-x_0-\Delta_0)^2}{2(\delta^2+\sigma^2)}\right] , \label{eq:marg1} \\ P(k) &=& \sqrt{\frac{2\delta^2}{\pi} } \exp \left[-2\delta^2(k-k_0)^2\right]. \label{eq:marg2} \earr Notice that the momentum distribution (\ref{eq:marg2}) is unaltered [$|\phi(k)|^2$ in (\ref{eq:gaussinv})]: obviously, the energy of the neutron does not change. Observe the additional spread in position $\delta'=(\delta^2+\sigma^2)^{1/2}$ and notice that the Wigner function and its marginals are always normalized to one. The uncertainty principle yields $\delta_k\delta' =\frac{1}{2} \sqrt{1+\sigma^2/\delta^2} > 1/2$. \section{Schr\"odinger-cat states in a fluctuating magnetic field} \label{sec-noise} \andy{noise} Let us now look in more detail at the experiment \cite{BRSW}. A polarized ($+y$) neutron enters a magnetic field, perpendicular to its spin, of intensity $B_0=0.28$mT, confined in a region of length $L=57$cm. Due to Zeeman splitting, the two neutron spin states travel with different speeds in the field. The average neutron wavenumber is $k_0=1.7 \cdot 10^{10}$m$^{-1}$ and its coherence length (defined by a chopper) is $\delta=1.1\cdot 10^{-10}$m. By travelling in the magnetic field, the two neutron spin states are separated by a distance $\Delta_0=2 m\mu B_0/\hbar^2 k_0= 16.1 \cdot 10^{-10}$m, one order of magnitude larger than $\delta$ (notice the factor 2, absent in the definition of the previous section). Observe that the neutron wave packet itself has a natural spread $\delta_t=(\delta^2+(\hbar t/2m\delta)^2)^{1/2} \simeq 15$cm (due to its free evolution for a time $t\simeq mL/\hbar k_0$); however, we shall neglect this additional effect, because it is irrelevant for the loss of quantum coherence. After the neutron has crossed the $B$-field only the $+y$ spin-component is observed and its Wigner function is readily computed \begin{eqnarray}}\newcommand{\earr}{\end{eqnarray} & & W(x,k)=\frac{1}{4\pi}\exp[-2\delta^2(k-k_0)^2]\nonumber\\ & & \times\left[\exp\left(-\frac{\left(x-\frac{\Delta}{2}\right)^2}{2\delta^2}\right) +\exp\left(-\frac{\left(x+\frac{\Delta}{2}\right)^2}{2\delta^2}\right) +2 \exp\left(-\frac{x^2}{2\delta^2}\right)\cos(k\Delta)\right]. \nonumber\\ \earr Notice that for $\Delta=0$ (no $B$-field) one obtains (\ref{eq:wiggauss}). Our interest is to investigate the loss of quantum coherence if the intensity of the $B$-field fluctuates, like in the previous section, yielding a random shift according to the law (\ref{eq:probdel}). In such a case, the average Wigner function reads \andy{Wm2} \begin{eqnarray}}\newcommand{\earr}{\end{eqnarray} & & W_{\rm m}(x,k)=\int d\Delta\;w(\Delta)\;W(x,k) = \frac{1}{4\pi}\exp[-2\delta^2(k-k_0)^2] \nonumber \\ & & \times \left[\sqrt{\frac{\delta^2}{\delta^2+\frac{\sigma^2}{4}}} \exp\left(-\frac{\left(x-\frac{\Delta_0}{2}\right)^2} {2\left(\delta^2+\frac{\sigma^2}{4}\right)}\right) +\sqrt{\frac{\delta^2}{\delta^2+\frac{\sigma^2}{4}}} \exp\left(-\frac{\left(x+\frac{\Delta_0}{2}\right)^2} {2\left(\delta^2+\frac{\sigma^2}{4}\right)}\right) \right. \nonumber\\ & &\quad\quad +\left. 2\exp\left(-\frac{x^2}{2\delta^2}\right) \exp\left(-\frac{\sigma^2 k^2}{2}\right)\cos(k\Delta_0)\right] \label{eq:Wm2} \earr \begin{figure} \centerline{\epsfig{file=noise.eps, width=\textwidth}} \caption{Fig. 2. Wigner function (\ref{eq:Wm2}) for $\Delta_0=0, k_0=1.7 \cdot 10^{10}$m$^{-1},\delta=1.1 \cdot 10^{-10}$m; $x$ is measured in units $10^{-10}$m and $p=k$ in units $10^{10}$m$^{-1}$. From above left to bottom right, $\sigma=0,0.5,1.0,1.5\cdot10^{-10}$m. Notice the asymmetry of the ``wiggles" around $k_0$ and their fragility at high values of momentum. Observe also the slight ``squashing" of the Gaussian components at high values of $\sigma$. } \end{figure} and the momentum distribution function yields \andy{Wm2pk} \begin{eqnarray}}\newcommand{\earr}{\end{eqnarray} P(k)= \sqrt{\frac{\delta^2}{2\pi}}\exp[-2\delta^2(k-k_0)^2] \left[ 1 + \exp\left(-\frac{\sigma^2 k^2}{2}\right)\cos(k\Delta_0)\right]. \label{eq:Wm2pk} \earr Notice also that, since only the $+y$-component of the neutron spin is observed, the normalization reads \begin{eqnarray}}\newcommand{\earr}{\end{eqnarray} N &=& \int dx\; dk\; W_{\rm m}(x,k) \nonumber\\ &=& \frac{1}{2} \left[1 +\sqrt{\frac{\delta^2}{\delta^2+\frac{\sigma^2}{4}}} \exp\left(-\frac{\Delta_0+4\delta^2\sigma^2 k_0^2} {8\left(\delta^2+\frac{\sigma^2}{4}\right)}\right) \cos\left(\frac{\delta^2}{\delta^2+\frac{\sigma^2}{4}} k_0\Delta_0\right)\right]. \earr Obviously, $N=1$ when no magnetic field is present ($\sigma=\Delta_0=0$). The Wigner function (\ref{eq:Wm2}) is plotted in Figure 2 for some values of $\sigma$. The off-diagonal part of the Wigner function (``trustee" of the interference effects) is very fragile at high values of momentum. This was already stressed in \cite{RSP,BRSW} and is apparent in the structure of the marginal distribution (\ref{eq:Wm2pk}): the term $\exp (-\sigma^2 k^2/2)$ strongly suppresses the interference effects at high $k$'s. \section{Decoherence parameter} \label{sec-dec} \andy{dec} One can give a quantitative estimate of the loss of quantum coherence by introducing a ``decoherence parameter," in the same spirit of Refs.\ \cite{MHS}. To this end, remember that the Wigner function can be expressed in terms of the density matrix $\rho$ as \andy{wigrho} \begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:wigrho} W(x,k) = \frac{1}{2\pi}\int d\xi\; e^{-ik\xi} \langle x+\xi/2|\rho|x-\xi/2\rangle , \eeq and that $\mbox{Tr}(\rho^2) =\mbox{Tr}\rho =1$ for a pure state, while $\mbox{Tr}(\rho^2) < \mbox{Tr} \rho=1$ for a mixture. Define therefore the {\em decoherence parameter} \andy{decpar} \begin{equation}}\newcommand{\eeq}{\end{equation}\label{eq:decpar} \varepsilon (\sigma) =1-\frac{\mbox{Tr}(\rho^2)}{(\mbox{Tr}\rho)^2} = 1 - \frac{2\pi \int dx\; dk\; W_{\rm m}(x,k)^2} {\left(\int dx\; dk\; W_{\rm m}(x,k)\right)^2} . \eeq This quantity is expected to vanish for $\sigma=0$ (no fluctuation of the $B$-field and quantum coherence perfectly preserved) and to become unity when $\sigma \to \infty$ (large fluctuations of the $B$-field and quantum coherence completely lost). Figure 3 confirms these expectations, that can also be proven analitically from (\ref{eq:Wm2}). \begin{figure} \centerline{\epsfig{file=epsilon.eps, width=\textwidth}} \caption{Fig. 3. Decoherence parameter. Left: $\epsilon$ as a function of $\delta$ and $\sigma$ (both in units $10^{-10}$m). Notice the peculiar behavior when $\delta>3$ and $1<\sigma<2$. Right: $\epsilon$ vs $\sigma$ (in $10^{-10}$m) for $\delta=1.1\cdot 10^{-10}$m (experimental value in \cite{BRSW}). } \end{figure} In Ref.\ \cite{BRSW}, $\delta=1.1 \cdot 10^{-10}$m and $\sigma$ is (presumably) very small, being the intensity of the $B$ field controlled with high accuracy. It is remarkable that the decoherence parameter is not a monotonic function of the noise $\sigma$, when $\delta>3\cdot10^{-10}$m and $1\cdot10^{-10}$m$<\sigma<2\cdot10^{-10}$m. This may be due to our very definition (\ref{eq:decpar}) or to some physical effect we do not yet understand. \medskip \noindent {\bf Acknowledgments:} We thank G.\ Badurek, H. Rauch and M. Suda for many useful discussions. This work was partially supported by the TMR Network ``Perfect Crystal Neutron Optics" (ERB-FMRX-CT96-0057) of the European Union.
\section{Introduction} \label{intr} The effect of pressure on the propagation of elastic waves in materials is essential for understanding interatomic interactions, mechanical stability of solids, phase transition mechanisms, material strength, and the internal structure of the Earth and other planets. However, little is known of the elasticity of solids at high pressure. The experimental study of the elasticity of materials under high pressure is challenging, as traditional methods have been applied only to moderate pressures. Ultrasonic measurements are generally limited to a few GPa \cite{jack90}, while Brillouin spectroscopy has been applied up to 25 GPa \cite{shim95}. We investigate the elasticity of three hexagonal transition metals at high pressure: iron, rhenium, and cobalt. High pressure properties of iron are of considerable geophysical interest as the Earth's solid inner core is composed primarily of this element. The elasticity of hcp iron is important for understanding the elastic anisotropy of the inner core \cite{creager,tromp,stix95}, and its super-rotation \cite{song}. Rhenium is the strongest metal known at high pressure \cite{jean91} and is widely used as a gasket material in diamond anvil cell experiments. We have chosen cobalt for this study because of its proximity to iron in the periodic table and as an example of a ferromagnetic hcp metal. All three of these metals have been studied experimentally under high pressure and their equations-of-state are well known. Iron transforms from the bcc phase at ambient conditions to hcp near 13 GPa \cite{jeph86}; the equation-of-state of the hcp phase has been measured up to 300 GPa \cite{mao91}. Recent advances in diamond anvil cell techniques have made it possible to evaluate the lattice strain in a polycrystal subjected to a non-hydrostatic stress field which can be associated with elastic constants. The elasticity of iron has been inferred by this method at high pressure (up to 210 GPa) \cite{singh98,mao98}. The equation-of-state of cobalt has been measured up to 80 GPa \cite{fuji96} and its elastic constants were obtained at zero pressure using traditional ultrasonic methods\cite{schober}. In the case of rhenium the equation-of-state is known to 215 GPa \cite{vohra87}, its elastic constants and their pressure derivatives have been ultrasonically measured at low pressure\cite{mangh74}. The same experimental method for evaluating lattice strains as in hcp iron has been applied to rhenium in the pressure range 18-37 GPa \cite{duffy98}. Iron has been studied widely with first-principles theoretical approaches because of its geophysical importance and the well known failure of the local density approximation (LDA) to the exchange-correlation potential to predict the ferromagnetic bcc ground state \cite{wang85}. This failure was a major impetus in the development of the generalized gradient approximation (GGA) \cite{PWI,PWII,PBE}. The equation-of-state of hcp iron under LDA and GGA is well known to high pressures\cite{bagno89,tera92,stix94,soed} and its elastic constants have been calculated by the full-potential linearized muffin-tin orbital method (FP-LMTO)\cite{soed}, and a total energy tight-binding (TB) method \cite{stix95,cohen97}. For hcp cobalt calculations have been performed with the LMTO method in the atomic sphere approximation for LDA \cite{min86} and the linearized combination of atomic orbital method (LCAO) for GGA \cite{leung91}. There is no previous theoretical work on the elastic constants of hcp cobalt. For rhenium only one study has focused on the hcp phase at high pressure \cite{fast95}; using FP-LMTO with LDA the equation-of-state and the elastic constants at zero pressure have been calculated. We organize the paper as follows. Section \ref{meth} elaborates the computational details of our first principles calculations and our approach to calculating the elastic constants, the elastic wave velocities, and the acoustic anisotropy. It is followed by a section presenting our results on the magnetic state of the materials studied, their $c/a$ ratios, the equation-of-state, and the elastic constants as functions of pressure. We compare our results in terms of the elastic wave velocities to high pressure experiments and the Earth's inner core. In section \ref{dis} we analyze the elastic anisotropy resulting from our calculations, recent experimental and theoretical results, and the predictions from a central nearest neighbor force model. Finally, we present our conclusions in section \ref{conc}. \section{Method} \label{meth} \subsection*{LAPW Total Energy Calculations} We investigate the energetics of hcp iron, cobalt, and rhenium using the full-potential linearized-augmented plane-wave method (LAPW) \cite{singhbook} with both LDA and GGA approximations to the exchange-correlation potential. For LDA the form of Hedin and Lundquist \cite{HL} and von Barth and Hedin \cite{BH} are used for non-magnetic and spin-polarized calculations, respectively. For GGA we adopt the efficient formulation of Perdew, Burke, and Ernzhofer \cite{PBE}. Core states are treated self-consistently using the full Dirac equation for the spherical part of the potential, while valence states are treated in a semirelativistic approximation neglecting spin-orbit coupling. We investigate ferromagnetic alignment in spin-polarized calculations for all metals and antiferromagnetism for iron. For consistency of the results all parameters in the calculations except for spin-polarization are kept fixed. For the $3d$ metals $3s$, $3p$, $3d$, $4s$, and $4p$ states are treated as valence electrons for all volumes. For rhenium we treat all electrons up to $4f$ as core, $5d$ and $6s$ as valence states. For rhenium we also have tested other configurations, such as including the $4f$ as valence states, which did not change our results significantly. The muffin-tin radii $R_{MT}$ are 2.0 Bohr for the $3d$ metals, and 2.3 Bohr for rhenium. As spin-orbit coupling of the valence electrons is important for the band structure and other properties of heavy elements, we consider the influence of the spin-orbit term on the equation-of-state for Re by including it in a variational step\cite{singhbook}. We carry out total energy calculations over a wide range of volumes for all three metals (0.7 $V_0$ - 1.2 $V_0$, with $V_0$ the zero pressure volume). At each volume we determine the equilibrium ratio of the lattice constants $c/a$ by performing calculations for several different values of this ratio. The equation-of-state is obtained by describing the energy-volume curve with a third order expansion in the Eulerian finite strain\cite{birch}. We have performed convergence tests with respect to Brillouin zone sampling and the size of the basis set, $R_{MT}K_{max}$, where $K_{max}$ is the largest reciprocal space wave-vector in the basis set. Converged results are achieved with a 12x12x12 special k-point mesh \cite{monk}, yielding 114 k-points in the irreducible wedge of the Brillouin zone for the hcp lattice, and up to 468 k-points for the monoclinic lattice used in elastic constants calculations. The number of k-points in the full Brillouin zone is well above the convergence criterion for elastic constant calculations established by Fast et al.\ \cite{fast95}. The size of the basis set is given by $R_{MT}K_{max}=9.0$, yielding 158 to 311 basis functions, depending on volume. Careful convergence tests show that with these parameters relative energies are converged to better than $0.1$ mRy/atom, magnetic moments to better than $0.05$ $\mu_{B}$/atom, and $c/a$ to within 0.025. \subsection*{Elastic Constants} We calculate the elastic constants as the second derivatives of the internal energy with respect to the strain tensor ($\varepsilon$). We choose the applied strains to be isochoric (volume-conserving) which has several important consequences: First, we assure the identity of our calculated elastic constants with the stress-strain coefficients, which are appropriate for the calculation of elastic wave velocities; this identity is non-trivial for finite applied pressure \cite{wall,BK65}. Second, the total energy depends on the volume much more strongly than on strain; by choosing volume conserving strains we obviate the separation of these two contributions to the total energy. Third, the change in the basis set associated with the applied strain is minimized, thereby minimizing computational uncertainties. We obtain the elastic constants at the equilibrium relaxed structure at any volume $V$ by straining the lattice, relaxing the symmetry allowed internal degrees of freedom, and evaluating the total energy changes due to the strain as a function of its magnitude $\delta$. The bulk modulus $K$ is calculated by differentiating the equation-of-state. For hexagonal crystals $K$ is the combination of elastic constants \begin{equation}\label{Kc} K=\left[C_{33}\left(C_{11}+C_{12}\right)-2 C_{13}^2\right]/C_S, \end{equation} with \begin{equation} C_S=C_{11}+C_{12}+2C_{33}-4C_{13}. \end{equation} The volume dependence of the optimized $c/a$ is related to the difference in the linear compressibilities along the $a$- and $c$-axes ($k_a$ and $k_c$). The dimensionless quantity $R$ describes this as \begin{equation} R=K(k_a-k_c)=-\frac{d \ln(c/a)}{d \ln V}. \end{equation} In terms of the elastic constants, \begin{equation} R=\left(C_{33}-C_{11}-C_{12}+C_{13}\right)/C_S. \end{equation} We calculate $C_S$ by varying the $c/a$ ratio at a given volume, according to the isochoric strain \begin{equation} \varepsilon(\delta)=\left(\begin{array}{ccc}\quad \delta \quad& \quad 0 & \quad 0 \\ \quad 0 \quad & \quad \delta & \quad 0 \\ \quad 0 \quad & \quad 0 & \quad \left(1+\delta\right)^{-2}-1 \end{array}\right). \end{equation} The corresponding energy change is \begin{equation} E\left(\delta\right)=E\left(0\right)+C_SV\delta^2+O(\delta^3). \end{equation} In the expressions for $C_S$, $K$, and $R$, $C_{11}$ and $C_{12}$ occur only as a sum. To separate these constants we determine their difference, $C_{11}-C_{12}=2C_{66}$ by applying an orthorhombic strain, space group {\it Cmcm}. For the strained lattice we use the two atom primitive unit cell, with the atoms in the Wyckoff position $4c$, coordinates ($y$,$-y$,1/4). The strain is \begin{equation} \varepsilon(\delta)=\left(\begin{array}{ccc} \quad \delta \quad & \quad 0 & \quad 0 \\ \quad 0 \quad & \quad -\delta & \quad 0 \\ \quad 0 \quad & \quad 0 & \quad \delta^2/\left( 1 - \delta^2 \right) \end{array}\right), \end{equation} leading to a change in total energy: \begin{equation} E\left(\delta\right)=E\left(0\right)+2C_{66}V\delta^2+O(\delta^4). \end{equation} In the unstrained lattice the atomic coordinate is $y=1/3$, but varies under strain\cite{nastar95}. We relax our calculations with respect to this internal degree of freedom. To determine $C_{44}$ we use a monoclinic strain, space group $C2/m$. The atomic positions in the two atom primitive unit cell are $(1/6,5/6,1/4)$. The strain applied \begin{equation} \varepsilon(\delta)=\left(\begin{array}{ccc} 0 & 0 & \delta \\ 0 & \delta^2/\left( 1-\delta^2\right) & 0 \\ \delta & 0 & 0 \end{array}\right) \end{equation} results in an energy change \begin{equation} E\left( \delta \right)=E\left(0\right)+2C_{44}V\delta^2+O(\delta^4). \end{equation} The equilibrium positions of the atoms are unaffected by this strain and do not need to be redetermined.\cite{nastar95} While for $C_{66}$ and $C_{44}$ the leading error term is of the order $\delta^4$, for $C_S$ it is of third order in $\delta$. It is therefore crucial to include positive and negative strains in the calculation for $C_S$. The strain amplitudes applied are typically nine values of $\delta$ covering $\pm$ 4\% for $C_S$; for $C_{66}$ and $C_{44}$, seven values of $\delta$ ranging to 6\% are applied. The elastic constants are then given by the quadratic coefficient of polynomial fits to the total energy results; the order of the polynomial fit is determined by a method outlined by Mehl \cite{mehl93}. From the full elastic constant tensor we can determine the shear modulus $\mu$ according to the Voigt-Reuss-Hill scheme\cite{hill63a} and hence the isotropically averaged aggregate velocities for compressional ($v_p$) and shear waves ($v_s$) \begin{equation} v_p=\sqrt{\left(K+\frac{4}{3}\mu\right)/\rho} \quad,\quad v_s=\sqrt{\mu/\rho}, \end{equation} with $\rho$ the density. More generally, the acoustic velocities are related to the elastic constants by the Christoffel equation \begin{equation} \left( C_{ijkl}n_j n_k-M\delta_{il}\right)u_i=0, \end{equation} where $C_{ijkl}$ is the fourth rank tensor description of elastic constants, $\mathbf{n}$ is the propagation direction, $\mathbf{u}$ the polarization vector, $M=\rho v^2$ is the modulus of propagation and $v$ the velocity. The acoustic anisotropy can be described as \begin{equation} \Delta_i=\frac{M_i[{\mathbf n_x}]}{M_i[100]}, \end{equation} where $\mathbf{n_x}$ is the extremal propagation direction other than $[100]$ and $i$ is the index for the three types of elastic waves (one longitudinal and the two polarizations of the shear wave). Solving the Cristoffel equation for the hexagonal lattice one can calculate the anisotropy of the compressional ($P$) wave as \begin{equation}\label{1} \Delta_P=\frac{C_{33}}{C_{11}}. \end{equation} For the shear waves the wave polarized perpendicular to the basal plane ($S1$) and the one polarized in the basal plane ($S2$) have the anisotropies \begin{equation}\label{2} \Delta_{S1}=\frac{C_{11}+C_{33}-2C_{13}}{4C_{44}} \quad,\quad \Delta_{S2}=\frac{C_{44}}{C_{66}}. \end{equation} While for $S2$- and $P$-waves the extremum occurs along the $c$-axis, for $S1$ it is at an angle of 45$^{\mbox{o}}$ from the $c$-axis in the $a$-$c$-plane. We note that an additional extremum may occur for the compressional wave propagation at intermediate directions depending on the values of the elastic constants. \section{results} \label{res} \subsection*{Magnetism} We find a stable ferromagnetic state only in cobalt. It is stabilized over a wide volume range with the magnitude of the moment decreasing with pressure in agreement with previous theoretical results on the pressure dependence of magnetic moments \cite{stix94} in other transition metals. Only at the smallest volume considered (50 Bohr$^3$, 180 GPa) is the moment vanishingly small (Fig.\ \ref{mm}). LDA and GGA yield consistent results and predict a zero pressure magnetic moment of 1.55 $\mu_B$, in excellent agreement with experiment (1.58 $\mu_B$ \cite{mey51}). In the case of hcp iron, we also investigate two antiferromagnetic states. The first consists of atomic layers of opposing spin perpendicular to the $c$-axis (afmI). The other arranges the planes of opposite spins normal to the $[100]$ direction in the hcp lattice; this can be described by the orthorhombic representation of the hcp unit cell (space group $Pmma$) with spin up in the $(1/4,0,1/3)$ and spin down in the $(1/4,1/2,5/6)$ position (afmII). We find that both structures are more stable than the non spin-polarized state and that afmII is energetically favored over afmI. For both antiferromagnetic states the moment is strongly pressure dependent. For afmI it vanishes at volumes larger than $V_0$ (Fig.\ \ref{mm}), in excellent agreement with results of Asada and Terakura \cite{tera92}. The other structure, afmII, possesses a magnetic moment well into the stable pressure regime of hcp iron, up to $\sim$ 40 GPa. (Fig.\ \ref{mm}). Because of frustration on the triangular lattice, it is possible that more complex spin arrangements such as incommensurate spin waves as for fcc iron \cite{uhl} or a spin glass are still more energetically favorable than afmII. Diamond anvil cell in situ M\"ossbauer measurements of hcp iron\cite{taylor91} have shown no evidence of magnetism in the hcp phase. The low antiferromagnetic moment we calculate in the stable hcp regime and the significant hysteresis of the bcc-hcp transition \cite{taylor91} might explain that no magnetism in hcp iron has been detected in the high pressure M\"ossbauer experiment. In this context it may be relevant that indirect evidence for magnetism exists at low pressure. Epitaxally grown iron-ruthenium superlattices have shown magnetism occurring in hcp iron multilayers\cite{maurer91}. Its character, however, is still controversial \cite{saint95,knab91}. \subsection*{$c/a$ Ratios} For all materials studied the $c/a$ ratio agrees with experimental data to within 2\% and is essentially independent of the exchange correlation potential (GGA or LDA). Equilibrium $c/a$ ratios for iron range from 1.58 at zero pressure to 1.595 at 320 GPa. This is consistent with experimental measurements\cite{jeph86,mao91} in the range of 15 to 300 GPa, which have shown considerable scatter. For cobalt the zero pressure $c/a$ ratio is calculated as 1.615, increasing to 1.62 at a pressure of almost 200 GPa. The zero pressure $c/a$ is slightly lower than the experimental value of 1.623 \cite{mey51}. Diamond anvil cell experiments have found a higher value of $c/a$, as much as the ideal value (1.633) \cite{fuji96}, this discrepancy might be due to the coexistence of hcp and metastable fcc cobalt in the polycrystalline sample \cite{fuji96}. The $c/a$ ratio for rhenium (1.615) does not change over the whole pressure range studied - and is in good agreement with experimental results (1.613)\cite{vohra87}. \subsection*{Equation-of-State} For the equation-of-state of rhenium, LDA shows better agreement with experimental data than does GGA (Fig.\ \ref{EOS}, Table \ref{EOS_table}). GGA overestimates the zero pressure volume and softens the bulk modulus, supporting a general pattern seen in prior density functional calculations using GGA for other $5d$ metals \cite{barb90,koer}. Including spin-orbit coupling in the calculation has little effect on the equation-of-state parameters, resulting in less than 1\% change in the zero pressure volume and 2\% in the bulk modulus. For cobalt, as for other $3d$ metals GGA is superior to LDA and reproduces the experimental equation-of-state to within 2\% in volume and 10\% in bulk modulus (Fig.\ \ref{EOS}, Table \ref{EOS_table}). The discrepancy in the equation-of-state parameters of hcp iron between non spin-polarized calculations and experiment is significantly larger than for the other two metals studied here (Table \ref{EOS_table}) or other transition metals \cite{barb90,koer}. The zero pressure volume is underestimated by $\sim$ 9\%, and the zero pressure bulk modulus is too stiff by 75\% (Table \ref{EOS_table}). Especially at low pressure the non-magnetic equation-of-state deviates considerably from experimental values, while at high pressure the agreement is very good (Fig.\ \ref{EOS}). The stabilization of antiferromagnetic states at low pressure can account for some of the discrepancy. For afmII magnetism persists to volumes smaller than $V_0$, resulting in a larger zero pressure volume, reducing the difference with experiment to 5\%, and lowering the bulk modulus considerably (Table \ref{EOS_table}). This is still larger than the difference in $V_0$ for cobalt and for cubic iron phases ($<$ 3\%) \cite{stix94,soed}. We attribute the remaining discrepancy between low pressure experimental data and the afmII equation-of-state (Fig.\ \ref{EOS}) to the approximations in GGA and the possible stabilization of more complex spin arrangements than those considered here. \subsection*{Elasticity} The agreement of the calculated elastic constants for cobalt and rhenium with zero pressure experimental results\cite{schober,mangh74} is excellent with a root mean square error of better than 20 GPa for both metals and both exchange-correlation potentials (Fig.\ \ref{cij}, Tables \ref{Co_cij} and \ref{Re_cij}). The initial pressure derivative of the elastic constants for rhenium is also well reproduced by the calculations (Fig.\ \ref{cij}). LDA and GGA exchange-correlation potentials give almost equally good agreement, the minor differences arising primarily from differences in the bulk modulus (Tables \ref{EOS_table}, \ref{Co_cij}, and \ref{Re_cij}). Our elastic constant calculations for rhenium and iron do not agree with the results of lattice strain experiments (Fig.\ \ref{cij}, Tables \ref{Re_cij} and \ref{Fe_cij}). For rhenium the overall agreement between these experiments and our elastic constants is better than for iron. $C_{11}$ and $C_{12}$ agree well over the pressure range of the experiments, while the other longitudinal ($C_{33}$) and off-diagonal constant ($C_{13}$) differ significantly (Fig.\ \ref{cij}, Table \ref{Re_cij})). The shear elastic modulus ($C_{44}$) shows the largest discrepancy of all elastic constants (factor of 1.5). For iron the results of the lattice strain experiments and our calculations are in reasonable agreement for the off-diagonal constants only. The longitudinal moduli we obtain at 60 Bohr$^3$ and 50 Bohr$^3$($\sim$ 50 GPa and $\sim$ 200 GPa, respectively) are larger by approximately 50\%. This is partly related to the overestimated bulk modulus in the calculations. The largest discrepancy, as in the case of rhenium, occurs in the shear elastic constants ($C_{44}$ and $C_{66}$). Aggregate properties such as the bulk and shear modulus, and the compressional and shear wave velocity are in somewhat better agreement between the theoretical results and the lattice strain experiment for both rhenium and iron (Figs.\ \ref{Re_v_p} and \ref{v_p}). For rhenium, theory and experiment differ by less than 15\% in bulk and shear modulus (Fig.\ \ref{Re_v_p}). For iron the discrepancy is considerable at intermediate pressure but becomes smaller with increasing pressure, as already seen for the equation-of-state (Figs.\ \ref{EOS} and \ref{v_p}). At $\sim$ 200 GPa the difference in bulk modulus between GGA and experiment is less than 5\% and the elastic wave velocities differ by $\sim$ 10\%. The shear modulus differs by 25\% even at high pressure. For iron the comparison with previous theoretical results gives a more coherent picture. While the longitudinal elastic constants from our calculations are larger by 10-20\% compared to TB\cite{cohen97} and FP-LMTO results \cite{soed} (Table \ref{Fe_cij}), the elastic anisotropy is similar: the pairs of longitudinal, shear, and off-diagonal elastic moduli display similar values. For the TB study this is true over the whole pressure range considered, for the FP-LMTO calculations only at low pressure; the ratio of the off-diagonal constants ($C_{12}/C_{13}$) is strongly pressure dependent in that study, varying from 0.9 at zero pressure to 0.6 at 400 GPa. \section{Discussion}\label{dis} We find that the elastic anisotropy (eqs.\ \ref{1} and \ref{2}) is similar for all three metals studied here. The magnitude of the anisotropy is 10$\pm$2\% for the longitudinal anisotropy and $\Delta_{S1}$, and 30$\pm$3\% for $\Delta_{S2}$ and is nearly independent of pressure (Fig.\ \ref{aniso}). This is consistent with the experimentally observed behavior of other hcp transition metals, all of which - except for the filled $d$-shell metals zinc and cadmium - show anisotropy of similar magnitude (Fig.\ \ref{aniso}). These results can be understood by comparison to a hcp crystal interacting with central nearest neighbor forces (CNNF) \cite{born}. For this model the elastic anisotropy is independent of the interatomic potential to lowest order in $P/C_{11}$, hence the anisotropy is dependent on the symmetry of the crystal only. Born and Huang \cite{born} have shown that from this CNNF model the elastic constants scale as 32:29:11:8:8 for $C_{33}$:$C_{11}$:$C_{12}$:$C_{13}$:$C_{44}$, yielding $\Delta_{P}=32/29$, $\Delta_{S1}=8/9$, and $\Delta_{S2}=45/32$ (Fig.\ \ref{aniso}). The experimentally determined elastic anisotropies of rhenium and hcp iron at high pressure from lattice-strain measurements differ substantially from our theoretical predictions, previous theoretical calculations, the behavior of all other hcp transition metals, and the simple CNNF model (Fig.\ \ref{aniso}). The shear anisotropy in particular is very different in the high pressure experiments as compared with all other relevant results. We suggest that this discrepancy may arise from assumptions made in the data analysis. In particular, the assumption that the state of stress on all crystallographic planes is identical\cite{singh98}. This condition may not be satisfied in a material undergoing anisotropic deformation (e.\ g.\ dominated by basal slip), behavior that is observed for many hcp transition metals. Theory shows much better agreement with lattice-strain experiments in terms of the isotropically averaged moduli. Even so, the agreement in the case of rhenium is much better than for iron. In this context it is important to point out that part of the discrepancy in the case of iron is due to the mutual inconsistency of the experimentally reported elastic constants and isotropic moduli. We have found that the elastic constants reported in Ref.\ \onlinecite{mao98} do not yield the values of $K$ and $\mu$ reported in the same paper. The reason for this discrepancy is unknown. \section{Conclusions} \label{conc} The equations-of-state and the elastic constant tensor at zero pressure and under compression for two ambient condition hcp transition metals, cobalt and rhenium, and for the high pressure phase of iron, hcp, are calculated by means of the first principles LAPW method. We find a ferromagnetic ground state for cobalt and an antiferromagnetic one for iron, with the antiferromagnetic moment vanishing at 60 Bohr$^3$. The equations-of-state for the metals are in good agreement with experiment, as are the elastic constants and pressure derivatives of the elastic constants for cobalt and rhenium at ambient pressure. Elastic constants for iron under high pressure as inferred from lattice-strain experiments differ significantly from our theoretical results. Similarly large discrepancies are also found between theory and high pressure static experiments on rhenium. The lattice-strain experiments also lead to large values of the shear anisotropy that differ from that of all other open shell hcp transition metals. Given the excellent agreement of the theoretical elastic constants for cobalt and rhenium with experiment at zero pressure, we suggest that a re-examination of the lattice-strain experiments for rhenium and iron is warranted. \section*{Acknowledgement} We greatly appreciate helpful discussions with Tom Duffy, Rus Hemley, Dave Mao, and Per S\"oderlind. The work was supported by the National Science Foundation under grant EAR-9614790, and by the Academic Stategic Alliances Program of the Accelerated Strategic Computing Initiative (ASCI/ASAP) under subcontract no B341492 of DOE contract W-7405-ENG-48 (REC). Computations were performed on the SGI Origin 2000 at the Department of Geological Sciences at the University of Michigan and the Cray J90/16-4069 at the Geophysical Laboratory, support by NSF grant EAR-9304624 and the Keck Foundation.
\section{Introduction} \label{section1} Effect of dissipation on quantum tunneling phenomenon lies out of the scope of the conventional weak coupling (Markovian) approximation, and has been studied extensively so far \cite{Caldeira83b,Leggett87}. In order to introduce dissipation, Caldeira and Leggett \cite{Caldeira81,Caldeira83a} involved a thermal reservoir which consists of oscillators influencing stochastic (Brownian) fluctuations to the tunneling two-level system. Their model, the so-called spin-boson model, is given by the following Hamiltonian, \begin{equation} {\cal H}=-\frac{\Delta}{2}\sigma^x+\sum_{i=1}^{N} \omega_i a^\dagger_i a_i +\frac{\sigma^z}{2} \sum_{i=1}^{N} f_i (a^\dagger_i+a_i), \label{Hamiltonian} \end{equation} where the operators $\{\sigma^\alpha\}$ denote the Pauli operators and the operators $a_i$ and $a_i^\dagger$ denote the bosonic annihilation and creation operators, respectively, for the $i$-th oscillation mode ($i=1 \sim N$). The set of these oscillators works as the above-mentioned reservoir with respect to the spin $1/2\mbox{\boldmath $\sigma$}$. The coupling coefficients $\{ f_i \}$ should be arranged so as to satisfy the so-called Ohmic-dissipation condition, \begin{eqnarray} J(\omega)&=&\pi\sum_i f^2_i \delta(\omega-\omega_i) \\ \label{ohmic_condition} &=& \left\{ \begin{array}{lr} 2\pi\alpha\omega & (\omega\ll\omega_{\rm c}) \\ 0 & (\omega\gg\omega_{\rm c}) \end{array} \right. . \end{eqnarray} The dimensionless constant $\alpha$ controls the strength of the dissipation, and the cut-off frequency $\omega_{\rm c}$ defines the energy unit throughout this paper; $\omega_{\rm c}=1$. The transverse field $\Delta$ induces quantum coherence (tunneling amplitude) between the spin up-and-down states, whereas the coupling to the reservoir disturbs the coherence. These conflicting effects are the central concern of the problem, which are apparently non-perturbative in nature. It is noteworthy that the spin-boson model (\ref{Hamiltonian}) is equivalent to the anisotropic Kondo model \cite{Chakravarty82,Bray82,Hakim84,Guinea85}. The parameter $\alpha$ controls the strength of the longitudinal Kondo coupling, whereas $\Delta$ controls the transverse-coupling strength. Thereby, the region $\alpha<1$ ($\alpha>1$) is identified as the the antiferromagnetic (ferromagnetic) Kondo phase; see the ground-state phase diagram shown in Fig. \ref{phase_diagram}. \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{phase_diagram.eps} \end{center} \caption{ The ground-state phase diagram of the spin-boson model ({\protect \ref{Hamiltonian}}) with the Ohmic dissipation. Quantitative estimation of the relaxation parameters such as the damping rate and the oscillation frequency is the main concern of this paper. } \label{phase_diagram} \end{figure} That is, in the region $\alpha<1$, the up-and-down spin states form a coherent (singlet) state through a certain tunneling amplitude, whereas in $\alpha>1$, the tunneling amplitude vanishes owing to the dissipation \cite{DeRaedt83,DeRaedt84}. Hence, in the region $\alpha>1$, the ground state is degenerated doubly. The effective tunneling amplitude is found to vanish at the phase boundary in the form \cite{Guinea85c,Guinea85b}, \begin{equation} \label{effective_tunneling} \Delta_{\rm eff}=\Delta \left( \frac{\Delta}{\omega_{\rm c}} \right)^{\frac{\alpha}{1-\alpha}} . \end{equation} Owing to the equivalence, we readily investigate the equilibrium (thermodynamic) properties by means of various theoretical techniques which have been developed so far for the Kondo problem. Note, however, the dynamical (non-equilibrium) properties are rather out of the scope of these analytical techniques; refer to the paper \cite{Lesage96} for a recent analytical approach. The dynamical properties, especially in the frequency range $\sim \Delta_{\rm eff}$, are the very concern of the present topic. The noninteracting-blip approximation \cite{Leggett87,Chakravarty84} was invented so as to describe the relaxation dynamics of this particular frequency range. This approximation is, however, justified in a rather limited parameter range $\alpha<0.5$. Hence, in order to investigate the relaxation properties, various numerical simulations have been performed so far. Chakravarty and Rudnick performed quantum Monte-Carlo simulation \cite{Chakravarty95}; the model they simulated is a one-dimensional long-range Ising model, which was derived \cite{Anderson70,Anderson71} through eliminating the reservoir (conduction election) degrees of freedom. They succeeded in obtaining the spectral function through the Pad\'e approximation of the imaginary time correlation function followed by the analytic continuation (Wick rotation). As a result, they found that the long-range asymptotic form of the dynamical correlation is governed by power law. This conclusion was supported by Strong \cite{Strong97} with use of the similar numerical technique. V\"olker \cite{Voelker98} followed the Chakravarty-Rudnick analysis, and reported that `quasi-particle' picture explains the spectral-function data. As a consequence, he obtained the damping rate and frequency. We utilize the picture to analyze our density-matrix-renormalization data. Costi and Kieffer \cite{Costi96,Costi97} used the Wilson numerical renormalization technique to calculate the spectral function. Their technique is particularly useful in order to investigate the Kondo fixed-point (very low temperature) physics definitely. Time-evolution simulation was performed with use of stochastic sampling (path integral) algorithm \cite{Stockburger98}. The method has an advantage over others that one can observe real-time dynamics directly. The stochastic-sampling error, however, grows as the evolution time is lengthened. In the present paper, for the first time, we perform Lanczos-diagonalization analysis of the spin-boson model (\ref{Hamiltonian}). In order to command vast assembly of the reservoir-oscillator modes, we adopted the density-matrix truncation scheme proposed by Zhang, Jeckelmann and White \cite{Zhang98}. (They studied the one-dimensional polaron system with this truncation technique.) The rest of this paper is organized as follows. In the next section, we propose an implementation of their algorithm to the spin-boson model (\ref{Hamiltonian}), and report the precision in detail. In Section \ref{section3}, by means of this new technique, we investigate the relaxation properties of the spin-boson model. We evaluate the dynamical susceptibility (resolvent), which is readily calculated in our scheme with use of the continued-fraction-expansion formula \cite{Gagliano87,Gagliano88}. Analyzing the analyticity of the resolvent, we confirm the above-mentioned `quasi particle' picture \cite{Voelker98} so as to obtain the damping rate and frequency. Our results are contrasted with the former quantum Monte-Carlo results \cite{Voelker98}. In the last section, we give summary and discussions. \section{ Application of the Hilbert-space-truncation algorithm to the spin-boson model } \label{section2} In this section, we propose a prescription for adopting the Hilbert-space-reduction method to the spin-boson model. We then demonstrate the precision of our new scheme. \subsection{Hilbert-space-reduction algorithm} \label{section2_1} Even a single oscillator (boson) spans infinite-dimensional Hilbert space. Therefore, in order to treat an oscillator with the exact diagonalization method, one must truncate the Hilbert space inevitably. In conventional simulations, the boson state is represented in terms of its occupation number, and those states whose occupation number exceeds a limit are disregarded (discarded). Zhang, Jeckelmann and White \cite{Zhang98} proposed an alternative representation and a truncation criterion. Applying their scheme to a polaron system (one-dimensional Holstein model), they demonstrated that the scheme yields very precise results, although the dimensionality of each oscillator is reduced to three. This truncation is called `(numerical) renormalization.' Their new bases are particularly efficient in those cases where the oscillator equilibrium position is shifted by a certain external force (coordinate-coordinate coupling). This is precisely the case of the present model (\ref{Hamiltonian}). In the following, we explain the detail how we adopt their algorithm to the spin-boson model. First, we limit the Hilbert-space dimensionality of each oscillator to $d$ dimensions except an oscillator with $D$ dimensions; see Fig. \ref{DMRG}. \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{DMRG.eps} \end{center} \caption{ Schematic drawing of the density-matrix-renormalization algorithm applied to the spin-boson model ({\protect \ref{Hamiltonian}}). } \label{DMRG} \end{figure} We call the $d$-dimensional oscillators `small oscillators,' and the $D$-dimensional one `big oscillator.' The choice of the big oscillator is to be made in sequence among the various reservoir oscillator modes ($\alpha=1 \sim N$). The sequence is continued until the relevant bases, explained below, become converged. In our experience, a few sweeps are sufficient for the convergence. The space of each small oscillator is spanned by the above-mentioned truncated relevant bases, and thus the creation and the annihilation operators should be represented in terms of these bases correspondingly. (Before renormalization, the space is allowed to be of the number representation ($|n\rangle$, $n=1 \sim d-1$), and the operators are expressed in terms of this subspace.) On the other hand, the space of the big oscillator is spanned by the occupation-number representation with $n=0 \sim D-1$. From this $D$-dimensional space, $d$ relevant bases are renormalized (extracted) in the following manner. Second, we diagonalize the Hamiltonian (\ref{Hamiltonian}) with the Lanczos technique to obtain the ground state $|\Psi_0 \rangle$. Note that the Hamiltonian consists of the operators whose matrix elements are already prepared in the above. The ground-state vector should be represented in the form, \begin{equation} | \Psi_0 \rangle = \sum_{i,j} \psi_{ij} |i\rangle_A |j\rangle_B, \end{equation} where the bases $\{ |i\rangle_A \}$ are of the big site, and the bases $\{ |j\rangle_B \}$ are of the remaining part of the system. Thereby, we construct the (reduced) density matrix subjected to the big oscillator, \begin{equation} \rho_{i i'}=\sum_j \psi_{i j}\psi^{*}_{i' j}. \end{equation} The new bases $\{ {\bf u}^\beta \}$ are determined so as to diagonalize the density matrix, \begin{equation} \rho {\bf u}^\beta = w_\beta {\bf u}^\beta. \end{equation} According to the proposal \cite{White92,White93}, the new relevant bases (subspace) should be spanned by the eigenvectors ${\bf u}^\beta$ up to the $d$-th largest weight (eigenvalue) $w_\beta$. That is, the new annihilation operator of the big oscillator is re-represented in terms of these $d$ relevant eigenvectors in the following manner, \begin{equation} [\tilde{b}_i]_{\beta \beta'}={}^{{\rm t}} {\bf u}^\beta b_i {\bf u}^{\beta'} . \end{equation} Because the renormalization is subjected to the reduced density matrix, the renormalization is called `density matrix' renormalization \cite{White92,White93}. The renormalization is continued for every oscillator modes ($i= 1\sim N$) successively until the relevant bases become fixed (converged). In our experience, a few sweeps are sufficient for the convergence. Finally, we mention how we chose the frequencies $\{ \omega_\alpha \}$ and the coupling constants $\{ f_i \}$. The chose is arbitrary under the constraint eq. (\ref{ohmic_condition}). We have uniformly distributed the oscillator frequency over the range $0<\omega<\omega_{\rm c} (=1)$, and determined the coupling constants so as to satisfy the constraint. We numbered the oscillator modes ($\alpha=1 \sim N$) in order of the frequency $\omega_\alpha$ (from the slowest to the fastest). \subsection{Precision of the Hilbert-space-truncation algorithm} In this subsection, we show the precision of the algorithm explained in the preceeding subsection. In Fig. \ref{shusoku_t5}, we plotted the transverse magnetization $\langle \sigma^x \rangle$ for the system $N=8$, $\Delta=0.3$ and $\alpha=0.5$ by means of the conventional truncation scheme; namely, the boson states are represented in term of the occupation number, and the occupation number is restricted within $n \le n_{\rm max}$. \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{shusoku_t5.eps} \end{center} \caption{ The transverse magnetization is plotted for the system with $N=8$, $\Delta=0.3$ and $\alpha=0.5$ with the dimensionality ${\cal D}$ of each oscillator varied. The plots $+$ are evaluated with the conventional occupation-number representation. The occupation number is restricted within $n \le n_{\rm max}$. Hence, ${\cal D}=n_{\rm max}+1$. The plots $\times$ are those evaluated with the density-matrix-truncation algorithm. The dimensionality of each small oscillator $d$ is varied (${\cal D}=d$) with the big oscillator dimensionality fixed ($D=6$). } \label{shusoku_t5} \end{figure} (Note that the transverse magnetization indicates a degree to what extent the tunneling is disturbed by the reservoir fluctuations.) Hence, the dimensionality of each oscillator is given by ${\cal D}=n_{\rm max}+1$. We observe that through increasing ${\cal D}$, the results converge to a certain limit. In the same plot, we show the results by means of the density-matrix-renormalization method for the same system as the above ($N=8$, $\Delta=0.3$ and $\alpha=0.5$). In this renormalization, we fixed the dimensionality of the big oscillator as $D=6$, and varied the small-oscillator dimensionality $d$; hence, ${\cal D}=d$. We see that, with fewer number of bases, the renormalization results converge more rapidly than the former occupation-number-% representation results. As is mentioned in the previous subsection, the equilibrium position of each oscillator is shifted by the coordinate-coordinate coupling. The relevant bases of the density-matrix renormalization are constructed so as to take into account of the fluctuations from the equilibrium position, whereas the occupation-number bases are rather inefficient to represent these shifted oscillator modes. In Fig. \ref{gosa_t5}, we show the relative error of the transverse magnetization of the density-matrix-% renormalization method ($D=6$); the error is defined as the deviation from the the conventional occupation-number-truncation diagonalization with ${\cal D}=n_{\rm max}+1=6$. The system parameters are the same as those shown in Fig. \ref{shusoku_t5}. \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{gosa_t5.eps} \end{center} \caption{ The relative error of the transverse magnetization with the density-matrix-renormalization method. We varied the number of remained bases $d$ with $D=6$ fixed. The system parameters are the same as those shown in Fig. {\protect \ref{shusoku_t5}}. } \label{gosa_t5} \end{figure} We observe, with very limited number of relevant bases ($d=2\sim3$), the density-matrix renormalization reproduces the full-diagonalization result precisely. The relative error is saturated to $\sim 10^{-7}$ due to the numerical round-off error for $d \ge 4$. The result indicates that a few relevant oscillator modes are of importance. In Fig. \ref{weight_t5}, we show the weight $w_\beta$ of each eigenstate ($i=1 \sim D$) for each reservoir oscillator ($\alpha=1 \sim N$). \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{weight_t5.eps} \end{center} \caption{ The weight $w^\beta$ (density-matrix eigenvalue) is plotted for each oscillator mode ($i=1\sim8$). } \label{weight_t5} \end{figure} The weight $w_\beta$ indicates the statistical weight of the state ${\bf u}^\beta$ contributing to the ground state. We notice that, in fact, the first few bases are particularly weighted, and the other states are irrelevant ($w_\beta<10^{-5}$), and to be ignored. Furthermore, it should be noted that these features are common to all the reservoir modes ($i=1 \sim N$). To summarize, in Fig. \ref{shusoku_t5}, we found that by means of the conventional occupation-number truncation, the result converges gradually as the occupation-number threshold $n_{\rm max}$ is increased. We see that the occupation numbers exceeding $n \approx 6$ are hardly excited. In the density-matrix renormalization, see Figs. \ref{gosa_t5}-\ref{weight_t5}, we found that only the first few states are particularly weighted. Hence, hereafter, we set the dimension of the big site to be $D=6$, and remain two relevant states ($d=2$) for each oscillator mode. We believe that the number of the reservoir oscillators is prior to the number of remained bases $d$ for each mode: Note that as the number of the reservoir modes is increased, the coupling constants $\{ f_i \}$ should be reduced correspondingly; the oscillators become less disturbed. In the following, we simulate the reservoir consisting of eighteen oscillator modes. Hence, the truncation error is expected to be reduced furthermore from those shown in this subsection. \section{ Density-matrix-renormalization analysis of the dissipative tunneling} \label{section3} With use of the method developed in the preceeding section, we investigate the relaxation properties in the ground state of the spin boson model (\ref{Hamiltonian}). These properties are extracted from the dynamical susceptibility, which is readily evaluated in our scheme. Our results of the relaxation coefficients are contrasted with those obtained by means of the quantum Monte-Carlo simulation \cite{Voelker98}. \subsection{Dynamic susceptibility --- continued-fraction-expansion formula} In this subsection, we evaluate the dynamical susceptibility, and compare ours with that obtained at an integrable point $\alpha=0.5$. The analyticity of the susceptibility is analyzed extensively in the next subsection so as to yield relaxation properties. Linear response theory states that the relaxation (non-equilibrium) process should be described in term of a certain ground-state equilibrium dynamical correlation function, unless the process is not so far from equilibrium. In other words, equilibrium dynamical correlation function does contain informations about non-equilibrium processes. For that purpose, we calculated the following dynamical susceptibility, \begin{eqnarray} \chi''(\omega)&=&\frac12 \int_{-\infty}^{\infty}{\rm d}t {\rm e}^{{\rm i}\omega t} \langle [\sigma^z(t),\sigma^z]\rangle \nonumber \\ \label{Czz} &=& {\rm Im}\left( \left\langle\sigma^z\frac{1}{\omega+E_{\rm g}-{\cal H}+{\rm i}\eta}\sigma^z \right\rangle \right. \nonumber \\ \label{dynamical_susceptibility} & & \ \ \ \ \ - \left. \left\langle\sigma^z\frac{1}{\omega-E_{\rm g}+{\cal H}-{\rm i}\eta}\sigma^z \right\rangle \right) \\ &=&{\rm Im}G(\omega) . \end{eqnarray} Some might wonder that the inverse matrix of the total Hamiltonian appearing in eq. (\ref{dynamical_susceptibility}) cannot be computed; this is true. The {\em expectation value} of the inverse of the Hamiltonian is, however, evaluated with use of the Gagliano-Balseiro continued-fraction formula \cite{Gagliano87,Gagliano88}, \begin{equation} \label{Gagliano_Balseiro} \left\langle f_0 \left| \frac{1}{z-{\cal H}} \right| f_0 \right\rangle = \frac{\langle f_0 | f_0 \rangle} { z-\alpha_0-\frac{\beta_1^2} { z-\alpha_1-\frac{\beta_2^2} { z-\alpha_2-\frac{\beta_3^2}{\ddots} } } }, \label{Gagliano-Balseiro} \end{equation} where the coefficients are given by the Lanczos tri-diagonal elements, \begin{eqnarray} |f_{i+1}\rangle &=& {\cal H}|f_i\rangle-\alpha_i|f_i\rangle -\beta_i^2|f_{i-1}\rangle, \\ \alpha_i &=& \langle f_i|{\cal H}|f_i\rangle/\langle f_i|f_i\rangle , \nonumber \\ \beta_i^2 &=& \langle f_i|f_i\rangle/\langle f_{i-1}|f_{i-1}\rangle \ \ \ (\beta_0 = 0). \nonumber \end{eqnarray} Therefore, through choosing of the Lanczos initial vector as $|f_0\rangle=\sigma^z |\Psi_0\rangle$, we readily evaluate the dynamical susceptibility by means of the same numerical procedure used in the preceeding Lanczos diagonalization. We expanded the formula (\ref{Gagliano-Balseiro}) up to the one-hundredth order. This is comparable with the iteration number needed to obtain the low-lying states in the Lanczos diagonalization. Therefore, the expansion is expected to yield precise result concerning low-lying frequencies. This frequency range is sufficient for our purpose, because we are concerned in the frequency range $\sim \Delta_{\rm eff}$. Moreover, in practice, at high frequencies, the magnitude of the spectral intensity is suppressed considerably. In Fig. \ref{S_t2_t5}, we plotted the spectral function, \begin{equation} \label{spectral_function} S(\omega)=\frac{\chi''(\omega)}{\omega} , \end{equation} for the parameters $\Delta=0.2$, $\alpha=0.5$ and $N=18$. (The spectral function is related closely to the linear-response function.) \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{S_t2_t5.eps} \end{center} \caption{ The spectral function $S(\omega)$ ({\protect \ref{spectral_function}}) is plotted for $\Delta=0.2$ and $\alpha=0.5$. The solid line is our density-matrix-renormalization result for $N=18$. The delta-function peaks are broadened into the Lorentz form with the width $\eta=0.022$. The dashed line shows a rigorous result ({\protect \ref{Toulouse}}), which is available at the Toulouse point $\alpha=0.5$. } \label{S_t2_t5} \end{figure} At the point $\alpha=0.5$, the Hamiltonian (\ref{Hamiltonian}) is reduced to being quadratic through a canonical transformation, and the spectral intensity is calculated exactly in the form \cite{Guinea85b}, \begin{eqnarray} \label{Toulouse} S(\omega) &=& \frac{8\tilde{\Delta}^2}{\pi} \frac{1}{\omega^2+4\tilde{\Delta}^2} \nonumber \\ & & \times \left( \frac{1}{\omega\tilde{\Delta}}\arctan\left(\frac{\omega}{\tilde{\Delta}}\right) + \frac{1}{\omega^2}\ln\left( 1+\frac{\omega^2}{\tilde{\Delta}^2} \right) \right), \end{eqnarray} ($\tilde\Delta=(\pi/4)\Delta^2/\omega_{\rm c}$). The rigorous result is plotted in Fig. \ref{S_t2_t5} as well. We observe that our numerical data reproduces the rigorous formula. The slight difference may be attributed to the spectral form (\ref{ohmic_condition}) of the reservoir around the cut-off frequency. We used a non-analytic reservoir spectrum which vanishes suddenly at the cut-off frequency. This difference may become irrelevant, if the tunneling amplitude is set to be sufficiently small compared with the cut-off frequency. \subsection{Analyticity of the dynamical susceptibility (resolvent) and the damping properties} \label{section3_2} In this subsection, we investigate the analyticity of the dynamical susceptibility (\ref{dynamical_susceptibility}), from which we extract informations about relaxation properties. In Fig. \ref{c_t2_t2}, we plotted the imaginary-time correlation function, \begin{eqnarray} \label{ondo_G} {\cal G}({\rm i}\omega) &=& - \int_0^\beta{\rm d}\tau \langle {\rm e}^{\tau{\cal H}}\sigma^z{\rm e}^{-\tau{\cal H}}\sigma^z \rangle_\beta {\rm e}^{{\rm i}\omega \tau} \\ &=& G({\rm i}\omega) \ \ \ \ (\beta\to\infty), \end{eqnarray} for the system with $\Delta=0.2$, $\alpha=0.2$ and $N=18$. \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{c_t2_t2.eps} \end{center} \caption{ The imaginary-time Green function ({\protect \ref{ondo_G}}) is plotted for $N=18$, $\Delta=0.2$ and $\alpha=0.2$. The solid line is our density-matrix-renormalization result. The plots $\diamond$ are those of the `quasi-particle' form ({\protect\ref{quasi_particle}}) with the damping coefficients $\lambda=0.052$ and $\tilde{\omega}=0.124$. } \label{c_t2_t2} \end{figure} As is shown in the plot, the numerical result is fitted well by the `quasi-particle' formula, \begin{equation} {\cal G}({\rm i}\omega)= \frac{A}{(\omega+\lambda)^2+\tilde{\omega}^2}. \label{quasi_particle} \end{equation} This fact was pointed out by V\"olker, who performed a quantum Monte-Carlo simulation; see Introduction. As is apparent from the definition (\ref{dynamical_susceptibility}), the (fitting) parameters $\lambda$ and $\tilde{\omega}$ give the damping rate and the frequency, respectively, of the `coordinate' $\sigma^z$ perturbed from the equilibrium position. According to the formula (\ref{quasi_particle}), the following relations hold; \begin{eqnarray} \label{lambda_omega} \lambda&=& \frac {\frac{{\rm d}}{{\rm d}\omega}\left({\cal G}({\rm i}\omega)\right)^{-1}} {\frac{{\rm d}^2}{{\rm d}\omega^2}\left({\cal G}({\rm i}\omega)\right)^{-1}} -\omega , \\ \label{omega_omega} \tilde{\omega}&=& \sqrt{ \frac{2 ({\cal G}({\rm i}\omega))^{-1}} {\frac{{\rm d}^2}{{\rm d}\omega^2} \left({\cal G}({\rm i}\omega)\right)^{-1}} -(\omega+\lambda)^2 . } \end{eqnarray} From these relations, we estimate the damping coefficients $\lambda$ and $\tilde{\omega}$. It is one of the advantages of our scheme that one can differentiate the function ${\cal G}({\rm i}\omega)$, because it is expanded in the analytic form (\ref{Gagliano_Balseiro}). in Fig. \ref{damp_t2_t2}, we plotted the right hand side of eqs. (\ref{lambda_omega}) and (\ref{omega_omega}); the system parameters are the same as those in Fig. \ref{c_t2_t2}. \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{damp_t2_t2.eps} \end{center} \caption{ The damping coefficients $\lambda$ ({\protect\ref{lambda_omega}}) and $\tilde{\omega}$ ({\protect\ref{omega_omega}}) are plotted; the system parameters are the same as those in Fig. {\protect \ref{c_t2_t2}}. Around the frequency range $\sim\Delta_{\rm eff}(\approx0.13)$, these quantities remain invariant, indicating that over this range of time, the dynamics is described by the quasi-particle picture ({\protect\ref{quasi_particle}}). } \label{damp_t2_t2} \end{figure} We see that these quantities are invariant actually over a certain frequency range $\sim \Delta_{\rm eff}(\approx0.13)$, and thus the quasi-particle picture (\ref{quasi_particle}) holds in the time range $\sim\Delta_{\rm eff}^{-1}$. As is explained in Introduction, we are concerned in the physics of this particular range of time. In consequence, we obtained the poles of $G(\omega)$ at $\omega=\pm \tilde{\omega} - {\rm i}\lambda$, which are {\em away} from the real axis. It is noteworthy that such irreversible features are not transparent in the original high-order-expansion result (\ref{Gagliano_Balseiro}). Through the above data analysis, such the relaxation features become clear; we discuss this point afterwards. In Fig. \ref{kekka}, we plotted the damping rate and the frequency. \begin{figure}[htbp] \begin{center}\leavevmode \epsfxsize=8.5cm \epsfbox{kekka.eps} \end{center} \caption{ The damping rate $\lambda$ and the frequency $\tilde{\omega}$ estimated by means of the density-matrix-renormalization algorithm; (a) $\Delta=0.1$, (b) $\Delta=0.2$ and (c) $\Delta=0.5$. These are to be contrasted with those of quantum Monte-Carlo method {\protect \cite{Voelker98}}. } \label{kekka} \end{figure} These damping coefficients are estimated both with the procedure as is shown in Fig. \ref{damp_t2_t2} and with the conventional least square fit by the function (\ref{quasi_particle}). Our results of the density-matrix renormalization confirm the former quantum Monte-Carlo study \cite{Voelker98}: The oscillation frequency $\tilde{\omega}$ is suppressed as the dissipation is strengthened. It vanishes around $\alpha\approx0.5$. This point indicates the transition between the under-damped and the over-damped oscillations. By means of the bosonization technique \cite{Guinea85}, this transition point was predicted to locate at $\alpha=0.5$ irrespective of the tunneling amplitude strength $\Delta$. Because in the bosonization technique, the band width (cut-off frequency) is supposed to be sufficiently large compared with the many-body interactions ($\Delta$ and $\alpha$), the validity in the strong-coupling region is not necessarily assured. In Fig. \ref{kekka}, in fact, we see that the $\tilde{\omega}$ plots become curved convexly, as the tunneling amplitude is strengthened, and accordingly it becomes evident that the transition point locates below $\alpha=0.5$. These features were reported in the paper \cite{Voelker98}, and were suspected to be due to a systematic numerical error caused by the critical slowing down. In our diagonalization calculation, we are free from the critical slowing down. Therefore, we conclude that for larger values of $\Delta$, the transition point actually shifts below $\alpha=0.5$. Furthermore, we confirm the report \cite{Voelker98} that at $\alpha=1/3$, the damping rate and the frequency coincide. At this point, the peaks around $\omega=\pm\tilde{\omega}$ of spectral intensity (\ref{spectral_function}) merge into a single peak so that this point was suspected to indicate a certain phase transition. The present result suggests that the point is merely the situation where the damping feature smears out the oscillatory behaviour, because these relative strengths exchange. We see that in Fig. \ref{kekka} (a) ($\Delta=0.1$), the plots for $\alpha>0.6$ are rather scattered (irregular). In that region, the quasi-particle poles shift towards the origin of the complex plane ($\pm\tilde{\omega}-\lambda\to0$), so that the dumping parameters degenerate into the slowest reservoir oscillation mode. In that situation, our method becomes inapplicable. Similar difficulty arises in the vicinity of the localization transition point $\alpha=1$ for larger values of $\Delta$. We are in a position to discuss the above quasi-particle feature furthermore in detail. It is noteworthy that the form (\ref{quasi_particle}) implies that the analyticity of the dynamical susceptibility is broken along the real axis; the upper and lower complex planes are not continued analytically. This is precisely due to the reservoir effect, which drives the spin state to be in equilibrium, violating the time-reversal symmetry. In our numerical result, however, the upper and the lower complex planes are continued analytically, although along the real axis, there distribute vast number of poles densely. This feature contradicts the above quasi-particle picture insisting isolated poles at $\omega=\pm\tilde\omega-{\rm i}\lambda$. This inconsistency is simply due to the fact that our reservoir spectrum is not continuous, and thus the time-reversal symmetry is maintained. Only through the limit of infinite oscillator modes, these poles merge into the `quasi-particle' poles away from the real axis. Hence, in order to extract relaxation properties, one should avoid the real axis, along which the result is suffered significantly from discontinuity of the reservoir modes. Along the imaginary axis, as is shown in Fig. \ref{c_t2_t2}, the result is smooth and monotonic, and is much easier to be fitted by the quasi-particle form. That is why we examined the imaginary-time Green function (\ref{ondo_G}) just as V\"olker did for analyzing his Monte-Carlo data computed at each Matsubara frequency. We stress that our resolvent is evaluated (expanded) in the analytical (continued-fraction) form as in eq. (\ref{Gagliano_Balseiro}). Therefore, as is shown previously in Fig. \ref{S_t2_t5}, one can obtain the spectral function (real-time Green function) immediately through performing the Wick rotation. This is one of the advantages of the present scheme over others. \section{Summary and discussions} \label{section4} We investigated the dissipative two-state system (\ref{Hamiltonian}) through applying the density-matrix Hilbert-space-truncation algorithm \cite{Zhang98}. An implementation of this algorithm is proposed, and the precision applied to the present model is reported in detail. We found that in our situation where the oscillator equilibrium position is shifted by the coordinate-coordinate coupling, the density-matrix renormalization works much better than the conventional occupation-number-truncation method. We have remained two relevant states for each oscillator mode, so that we succeeded in treating eighteen oscillators, keeping the truncation error considerably small. We belive that the number of tractable oscillators is more crucial in observing the relaxation properties than the `quality' (fidelity) of each oscillator mode. In fact, we reproduced the dynamical susceptibility at the Toulouse point $\alpha=0.5$, at which a rigorous formula is known. We stress that, in our scheme, the susceptibility is evaluated in the {\em analytical} continued-fraction form (\ref{Gagliano_Balseiro}), which is apparently advantageous over others in performing the analytic continuation (Wick rotation); in order to observe relaxation (time irreversible) properties from numerical data, which apparently possesses the time-reversal symmetry, we need to examine the analyticity of the resolvent (dynamical susceptibility); see Section \ref{section3_2}. In fact, after the immediate analytic continuation to the imaginary axis, we found that the result is well fitted by the form (\ref{quasi_particle}), confirming V\"olker's finding \cite{Voelker98} with quantum Monte-Carlo method. That is, the susceptibility possesses the so-called quasi-particle poles away from the real axis $\omega=\pm\tilde{\omega}-{\rm i}\lambda$. In consequence, we obtained the damping rate $\lambda$ and the frequency $\tilde{\omega}$. Both agree with those of the quantum Monte-Carlo method \cite{Voelker98}. In particular, our results confirm the former report that for larger values of $\Delta$, the transition point between the over-damped and the under-damped oscillations shifts downwards ($\alpha_{\rm c}<0.5$). For studying such critical property, our method is advantageous over the Monte-Carlo method, because our method is free from the critical slowing down. \section*{Acknowledgement} Our program is based on the subroutine package TITPACK ver. 2 coded by professor H. Nishimori. \input{biblio} \end{document} E-mail: <EMAIL> Tel: +81-86-252-7809 Fax: +81-86-252-7830 @
\section{Introduction} In the 1700's, Immanuel Kant and the Marquis de Laplace proposed that the solar system collapsed from a gaseous medium of roughly uniform density (\cite{kan55}, \cite{lap96}). A flattened gaseous disk -- the protosolar nebula -- formed out of this cloud. The Sun contracted out of material at the center of the disk, while the planets condensed in the outer portions. Despite its simplicity, this model suffered from the {\it angular momentum problem} inherent to star formation: a cloud of gas and dust with the diameter of the solar system and the mass of the Sun has too much angular momentum to collapse to the Sun's present size. It took 150 years to solve this problem. C. von Weisz\"acker worked out the basic physics of a viscous accretion disk, building on previous realizations that a turbulent viscosity could move material inwards and angular momentum outwards through the protosolar nebula (\cite{von43}, \cite{von48}). With L\"ust's steady-state solution (\cite{lus52}), Kant's nebular hypothesis no longer suffered from the angular momentum problem and became the leading model for solar system formation. Disk accretion is now known to power a wide range of astronomical sources. Starting with Kuiper's landmark study of $\beta$ Lyr, viscous accretion disks have become central to our understanding of interacting binaries, where one star fills its inner Lagrangian surface and transfers matter into a disk surrounding a smaller companion (\cite{kui41}). A disk surrounding a supermassive black hole is widely believed to drive the phenomena observed in active galactic nuclei (e.g., \cite{ulr97}). Accretion disks remain an important issue in star and planet formation, as described throughout this volume. One striking feature of recent observations is the common phenomena observed in different accreting systems. All disks vary in brightness. These fluctuations range from rapid flickering on time scales as short as a few milliseconds up to long-lived eruptions with durations of decades or centuries. Most accreting systems lose mass; practically all disks that lose mass eject material in a well-collimated jet (\cite{liv97}). These common phenomena occur in systems with luminosities ranging from much less than 1 \ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi~in some interacting binaries up to $10^{11}~\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi$ in active galactic nuclei. \vskip 2ex \epsffile{fig1.eps} \footnotesize \baselineskip=10pt \vskip -4ex \noindent Figure 1 - Light curves for SS Cyg, a cataclysmic binary; CI Cyg, a symbiotic binary; and V1057 Cyg, an FU Ori variable star. Despite a different size scale in each system, the light curves have common features that indicate a similarity in their underlying physics. \vskip 4ex \normalsize \baselineskip=12pt As an example of these common phenomena, Figure 1 shows light curves for SS Cyg -- a cataclysmic binary with an orbital period of $P_{orb}$ = 6.6 hr (\cite{can98}); CI Cyg -- a symbiotic binary with $P_{orb}$ = 855 days (\cite{ken86}, \cite{bel92}); and V1057 Cyg -- an FU Ori variable that is apparently a single star (\cite{ken91a}). The 3--5 mag eruptions of these systems occur when the accretion rate through the disk increases. Aside from the deep eclipses in CI Cyg, the three light curves have common features, with a rapid rise, a brief interval at maximum, and a long decay. All three systems resemble A- or F-type stars at maximum light. This spectrum cools as the brightness fades. Substantial mass loss is associated with each system; the mass loss probably increases with the rise in brightness and decreases as the brightness declines. These similarities suggest that the same physical processes govern the evolution of disks ranging in size from 0.1 \ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi~to several tens of AU. \vskip -2ex \begin{table}[htb] \begin{center} \caption{Types of disk systems} \begin{tabular}{lccc} \hline System & Distance (pc) & Angular size (arcsec) & $N_{res}$ \\ \hline BH + MS & 1000 & $\mathrel{\copy\simlessbox} 10^{-5}$ & 100 \\ NS + MS & 1000 & $\mathrel{\copy\simlessbox} 10^{-5}$ & 100 \\ WD + MS & 100~ & $\mathrel{\copy\simlessbox} 10^{-3}$ & 100+ \\ MS + SG & 100~ & $\mathrel{\copy\simlessbox} 10^{-3}$ & 10+ \\ MS + RG & 1000 & $\mathrel{\copy\simlessbox} 10^{-3}$ & 10--100 \\ \\ $\beta$ Pic & 10+ & 1--10 & 10--100 \\ Young star & 100+ & 1--2 & 10--20 \\ \\ AGN & $10^8$ & $\mathrel{\copy\simlessbox} 10^{-3}$ & 1--10 \\ \hline \end{tabular} \end{center} \end{table} The common features of accreting systems provide good tools to test physical models of disks. Table 1 summarizes the main types of disk-accreting systems known today and provides simple comparisons of their angular sizes and the number of resolution elements across the face of the disk, $N_{res}$, that can be derived from modern observations. The first set of systems includes (i) compact binaries containing a disk surrounding a black hole (BH), a neutron star (NS), or a white dwarf star (WD), and (ii) wider binaries with a disk surrounding a main sequence star. A low mass main sequence star usually feeds the disk in compact binaries; a more evolved subgiant or red giant star feeds the disk in the wider binaries. In both types of binaries, the angular size of the disk is small compared to the resolution possible with current ground-based or space-based observatories. However, eclipses of the disk by the secondary star can be used to divide the disk into many resolution elements and to map out the physical structure of the disk with a few simplifying assumptions (\cite{ho85a}, \cite{ho85b}). The dusty disks in the second group of systems are large enough to image directly at optical and radio wavelengths, although few systems have been mapped in much detail (\cite{smi84}, \cite{jay98}). The final group of active galactic nuclei have only recently been mapped with VLBI techniques; $N_{res}$ will probably increase as these techniques become more sophisticated (\cite{her98}). \section{Steady Disks} Accretion disks work to transport mass radially inwards and angular momentum radially outwards. To understand how a disk manages this feat, consider a thin ring with two adjacent annuli at distances $r_1$ and $r_2$ from a central star (Figure 2). Material in these annuli orbits the central star with velocities, $v_1$ and $v_2$. The velocity difference between the two annuli, $v_1 - v_2 > 0$, produces a frictional force that attempts to equalize the two orbital velocities. The energy lost to friction heats the annuli; some disk material then moves inwards to conserve total energy. This inward mass motion increases the angular momentum of the ring; some disk material moves outwards to conserve angular momentum. Energy and angular momentum conservation thus lead to an expansion of the ring in response to frictional heating. The ring eventually expands into a disk, which generates heat (and radiation) at a level set by the accretion rate. \vskip -1ex \epsfxsize=4.0in \epsffile{fig2.eps} \footnotesize \baselineskip=10pt \vskip -3ex \noindent{Figure 2 - Schematic view of two adjacent annuli in a disk surrounding a compact star. Annulus `1' lies inside annulus `2' and orbits the star (filled circle) at a higher velocity, $v_1 > v_2$.} \baselineskip=12pt \normalsize \subsection{Accretion Luminosities and Temperatures} Figure 3 shows the standard disk geometry. In a {\it steady} disk, material drifts radially inward at a constant rate, $\mdot$. For an infinite disk, the total luminosity generated by accretion is $L_{acc}$ = $G \ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot/\ifmmode {R_{\star}}\else $R_{\star}$\fi$, which is \begin{equation} L_{acc} = 314~\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi \left ( \frac{\ifmmode {M_{\star}}\else $M_{\star}$\fi}{1~\ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi} \right ) \left ( \frac{\mdot}{10^{-5}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi} \right ) \left ( \frac{1~\ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi}{\ifmmode {R_{\star}}\else $R_{\star}$\fi} \right ) \end{equation} \epsfxsize=5.in \epsffile{fig3.eps} \footnotesize \baselineskip=10pt \vskip -23ex \noindent{Figure 3 - Schematic view of a disk that extends to the stellar photosphere. Disk material drifts radially inward until it reaches the boundary layer, where the rotational velocity slows to match the stellar rotational velocity.} \baselineskip=12pt \normalsize \vskip 3ex \noindent in familiar units. As material drifts inwards, half of this accretion energy is radiated by the disk, \begin{equation} L_{disk} = \frac{1}{2} L_{acc} ~ . \end{equation} \noindent The other half of the accretion energy becomes kinetic energy of orbital motion around the central star. In most cases, disk material drifts inwards until it reaches the stellar photosphere (Figure 3). At this point, disk material with an angular velocity of $\Omega = (G \ifmmode {M_{\star}}\else $M_{\star}$\fi/R_{\star}^3)^{1/2}$ must slow down to match the stellar angular velocity, $\Omega_{\star}$. This transition occurs in the ``boundary layer,'' a narrow ring with a radial size, \begin{equation} R_{bl,d} \sim h_{bl}^2/\ifmmode {R_{\star}}\else $R_{\star}$\fi \ll \ifmmode {R_{\star}}\else $R_{\star}$\fi, \end{equation} \noindent where $h_{bl}$ is the local scale height ($h_{bl}/\ifmmode {R_{\star}}\else $R_{\star}$\fi \ll 1$; \cite{pri77}, \cite{pri79}, \cite{reg83}, \cite{pap86}, \cite{kle87}, \cite{reg88}, \cite{kle91}). The energy lost in the boundary layer is the difference between the rotational energy per unit mass in the disk and the rotational energy per unit mass in the central star: \begin{equation} L_{BL} = \frac{1}{2} L_{acc} ~ \left ( \frac{\omega_{disk}^2 - \omega_{\star}^2}{\omega_{disk}^2} \right ) ~ . \end{equation} In some accreting systems, the magnetic field of the central star is strong enough to truncate the disk at a radius, $R_m > \ifmmode {R_{\star}}\else $R_{\star}$\fi$. A rough estimate of the truncation radius is the Alfv\'en radius, \epsfxsize=5.in \epsffile{fig4.eps} \footnotesize \baselineskip=10pt \vskip -15ex \noindent{Figure 4 - Schematic view of a disk truncated by a strong dipolar magnetic field. The inner radius of the disk lies close to the corotation radius, where the disk and the stellar photosphere have the same angular velocity. Material falls from the inner disk onto the star along the magnetic field lines and produces two circular rings surrounding the magnetic axis.} \vskip 4ex \normalsize \baselineskip=12pt \begin{equation} R_m \approx R_A \approx \left ( \frac{\mu_{\star}^4}{2 G \ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot^2} \right )^{1/7} \end{equation} \noindent where $\mu_{\star}$ is the magnetic dipole moment of the central star. More rigorous estimates of $R_m$ depend on the magnetic field geometry (\cite{lip78}, \cite{gho79}, \cite{lip80}, \cite{cam87}, \cite{kon91}, \cite{yi94}, \cite{arm95}, \cite{li96}, \cite{wan97}). The disk then has a total luminosity \begin{equation} L_{disk} = \frac{1}{2} L_{acc} ~ \left ( \frac{\ifmmode {R_{\star}}\else $R_{\star}$\fi}{R_m} \right ) \end{equation} \noindent In this geometry, disk material falls onto the star along magnetic field lines (Figure 4). The infalling gas shocks and produces two rings at stellar latitudes, $\pm b$, if the magnetic and rotational axes are parallel (\cite{gho79}, \cite{kon91}). These rings are compressed into tilted ellipses if the magnetic axis is inclined with respect to the rotational axis, as expected in most systems (\cite{ken94}, \cite{mah98}). In both cases, the luminosity of one ring is \begin{equation} L_{ring} = \frac{1}{2} L_{acc} ~ \left ( 1 - \frac{1}{2} \frac{\ifmmode {R_{\star}}\else $R_{\star}$\fi}{R_{mag}} \right ) \end{equation} To derive the temperature structure of the disk and the boundary layer or magnetic accretion ring, we again use conservation of energy. The disk temperature assumes an approximate balance between the energy lost from blackbody radiation and the torque due to frictional heating from mass motion through a distance $\Delta R$ in a gravitational potential: \begin{equation} G \ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot \left ( \frac{1}{R} - \frac{1}{R + \Delta R} \right ) \approx 2 \cdot 2 \pi R \Delta R ~ \sigma T_d^4 \end{equation} \noindent For $\Delta R \ll R$, this expression is \begin{equation} G \ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot \frac{\Delta R}{R^2} \approx 4 \pi R \Delta R \sigma T_d^4 \end{equation} \noindent which results in \begin{equation} T_d \approx \left ( \frac{G \ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot}{4 \pi \sigma R^3} \right )^{1/4} \propto R^{-3/4} \end{equation} The exact temperature of a steady disk depends on how the torque transports energy through the disk. Standard models assume that frictional heating vanishes at the stellar photosphere, which leads to an extra term in the energy equation (8) above (\cite{lyn74}, \cite{sha73}). The temperature is then (\cite{bat74}): \begin{equation} T_d(R) = T_{acc} \left ( \frac{R}{\ifmmode {R_{\star}}\else $R_{\star}$\fi} \right )^{-3/4} \left ( 1 - \sqrt{\ifmmode {R_{\star}}\else $R_{\star}$\fi/r} \right )^{1/4} \end{equation} \noindent where \begin{equation} T_{acc} = 13000~{\rm K} \left ( \frac{\ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot}{10^{-5}~\ifmmode {\rm M_{\odot}^2\,yr^{-1}}\else $\rm M_{\odot}^2\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi} \right )^{1/4} \left ( \frac{1~\ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi}{\ifmmode {R_{\star}}\else $R_{\star}$\fi} \right )^{3/4} . \end{equation} \noindent This expression for the temperature vanishes at the stellar surface, because the frictional heating vanishes. The temperature reaches a maximum at $R/\ifmmode {R_{\star}}\else $R_{\star}$\fi$ = 49/36 (\cite{bat74}): \begin{equation} T_{max} = 6500~{\rm K} \left ( \frac{\ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot}{10^{-5}~\ifmmode {\rm M_{\odot}^2\,yr^{-1}}\else $\rm M_{\odot}^2\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi} \right )^{1/4} \left ( \frac{1~\ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi}{\ifmmode {R_{\star}}\else $R_{\star}$\fi} \right )^{3/4} \end{equation} \noindent Different boundary conditions at the stellar surface yield slightly different, non-analytic, temperature laws (\cite{pop91a}, \cite{pop91b}). These results are usually close to the standard temperature law, equation (11). The temperature of the boundary layer depends on the mass accretion rate. At high \mdot, the region is optically thick (\cite{lyn74}, \cite{tyl81}). Energy generated in the thin ``dynamical boundary layer'' (equation 3) should diffuse over a broader ring with a size comparable to the thermal scale height, $h_{\rm bl}$ (\cite{pap86}). The scale height is smaller than the stellar radius in most cases. This ``thermal boundary layer'' is then much larger than the dynamical boundary layer and has a temperature 4--5 times larger than the maximum disk temperature (\cite{ken84}). For low $\mdot$, the boundary layer is optically thin and cannot radiate energy efficiently ({\cite{pri77}, \cite{pri79}). The thermal scale height is then comparable to the stellar radius, and the boundary layer resembles a hot stellar corona (\cite{pri77}, \cite{pri79}). The details of boundary layer structure are sensitive to properties of the stellar photosphere and the viscosity mechanism, but these general considerations hold for many physical settings (\cite{ber92}, \cite{ber93}, \cite{pop93}, \cite{hu95a}, \cite{hu95b}, \cite{god95}, \cite{god96}, \cite{pop96}, \cite{ida96}, \cite{kle96}, \cite{kle97}). The temperature of magnetic accretion rings also depends on geometry. If the magnetic axis is aligned with the rotational axis, the rings have equal luminosities and temperatures. The temperature is also constant with azimuth along each ring (\cite{mah98}). The two rings have unequal luminosities when the magnetic axis is tilted with respect to the rotational axis, as needed to produce the observed light variations in magnetic cataclysmic variables and T Tauri stars (\cite{cro90}, \cite{ken94}). The luminosity also varies with azimuth along each ring (\cite{mah98}). In the aligned case, the infalling matter shocks above the stellar surface and forms identical optically thick accretion columns that radiate in the Balmer and Paschen continua (\cite{lam98}, \cite{cal99}). The surface area of this emission is comparable to the surface area of the boundary layer, so the temperature is still 3--5 times larger than the maximum disk temperature (\cite{cal99}). The accretion columns are probably quite different in non-aligned cases, but their structure has not been addressed. \vskip -2ex \begin{table}[htb] \begin{center} \caption{Accretion and nuclear energies} \begin{tabular}{lcc} \hline Object & $\epsilon_g$ (\ifmmode {\rm erg~g^{-1}}\else $\rm erg~g^{-1}$\fi) & $\epsilon_{nuc}$ (\ifmmode {\rm erg~g^{-1}}\else $\rm erg~g^{-1}$\fi) \\ \hline T Tauri star & $5 ~ \times ~ 10^{14}$ & $5 \times 10^{13}$ \\ main sequence star & $2 ~ \times ~ 10^{15}$ & $4 \times 10^{18}$ \\ white dwarf star & $1 ~ \times ~ 10^{17}$ & $4 \times 10^{18}$ \\ neutron star & $1 ~ \times ~ 10^{20}$ & $6 \times 10^{18}$ \\ \hline \end{tabular} \end{center} \end{table} To judge the effectiveness of accretion as an energy source, Table 2 compares the gravitational potential energy, $\epsilon_g = G \ifmmode {M_{\star}}\else $M_{\star}$\fi/\ifmmode {R_{\star}}\else $R_{\star}$\fi$, of various 1~\ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~stars with the nuclear energy generation rate of the dominant fusion source, $\epsilon_{nuc}$. If a star can accrete {\it and burn} material from a disk, then the ratio of the accretion luminosity to the nuclear luminosity is simply the ratio of these two quantities, $r \equiv L_{acc}/L_{nuc}$ = $\epsilon_g/\epsilon_{nuc}$. This expression ignores the difference in accretion and nuclear time scales, but it provides a simple guide to the importance of accretion as a long-term energy source. Table 2 shows that accretion is rarely the major energy source for main sequence stars and white dwarfs; nuclear energy will always overwhelm accretion by factors of 40--2000 on long time scales. Accretion is an excellent energy source for pre-main sequence stars and neutron stars. It is roughly 10 times more effective than deuterium burning in pre-main sequence stars and nearly 20 times more productive than complete hydrogen burning in neutron stars. \subsection{Turbulent Viscosity and Disk Timescales} The source of the frictional heating in accretion disks remains controversial. Ordinary molecular viscosity is too small to generate mass motion on a reasonable time scale, $\sim$ days for disks in interacting binary systems and $\sim$ years to decades for disks in pre-main sequence stars. The large shear between adjacent disk annuli suggests that disks might be unstable to turbulent motions, which has led to many turbulent viscosity mechanisms. Convective eddies, gravitational instabilities, internal shocks, magnetic stresses, sound waves, spiral density waves, and tidal forces have all been popular turbulence mechanisms in the past three decades (see, for example, \cite{lyn74}, \cite{sha73}, \cite{cab87}, \cite{lin87}, \cite{vis89}, \cite{bal91}, \cite{haw91}, \cite{tou92}, \cite{sto96}, \cite{cab96}, \cite{vis96}, \cite{roz96}, \cite{reg97}, \cite{arm98}). Recent work has shown that magnetic stresses in a differentially rotating disk inevitably lead to turbulence (\cite{bal96}, \cite{bal98}, \cite{lin95}). The growth time and effectiveness of these magnetohydrodynamic mechanisms make them the current leading candidate for viscosity in most applications. How this turbulence leads to significant mass motion in a real accretion disk remains an unsolved problem. Shakura \& Sunyaev side-stepped the basic uncertainties of viscosity mechanisms when they developed the popular ``$\alpha$-disk'' model (\cite{sha73}; see also \cite{von48}). In this approach, the frictional heating in the disk is due to a turbulent viscosity, $\nu = \alpha c_s h$, where $c_s$ is the sound speed, $h$ is the local scale height of the gas, and $\alpha$ is a dimensionless constant. This concept is similar to the mixing length theory of convection, with $\alpha$ serving the role of the mixing length. In most applications, $\alpha \mathrel{\copy\simlessbox}$ 1--10; $\alpha$ needs to exceed $\sim 10^{-4}$ to allow material to move inwards on a reasonable time scale. This viscosity definition orders the important time scales for the disk (\cite{lyn74}). The shortest disk time scale is the dynamical (orbital) time scale, which increases radially outward: \begin{equation} \tau_D \approx {\rm 0.1~day}~ \left ( \frac{R}{1~\ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi} \right )^{3/2} \left ( \frac{1~\ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi}{\ifmmode {M_{\star}}\else $M_{\star}$\fi} \right )^{1/2} \end{equation} \noindent The thermal time scale measures the rate that energy diffuses through the disk, \begin{equation} \tau_T \approx {\rm 1~day}~ \left ( \frac{R}{1~\ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi} \right )^{11/8} ~ . \end{equation} \noindent The thermal time scale is intermediate between the dynamical time scale and the viscous time scale, which measures the rate matter diffuses through the disk, \epsfxsize=5.5in \hskip 12ex \epsffile{fig5.eps} \footnotesize \baselineskip=10pt \vskip -4ex \noindent{Figure 5 - Schematic evolution of a point-like mass enhancement, a blob, in a viscous accretion disk. A central star lies at the center of the disk with an outer edge indicated by the dashed circle. The disk initially has a smooth radial decrease in the effective temperature, T, and the surface density, $\Sigma$ (panel A). Differential rotation between adjacent annuli smooths the blob into a ring on the orbital time scale, $\tau_D$, which produces $\delta$-function increases in T and $\Sigma$ (panel B). The thermal energy in the blob moves inwards and outwards on the thermal time scale, $\tau_T$, which produces a gaussian-like perturbation in the temperature (panel C). The mass motion on this time scale is small; $\Sigma$ changes little on the thermal time scale. The mass perturbation spreads out through the disk on the viscous time scale, $\tau_V$, as indicated in panel D.} \vskip 4ex \normalsize \baselineskip=12pt \begin{equation} \tau_V \approx \frac{\rm 40~day}{\alpha} ~ \left ( \frac{R}{1~\ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi} \right )^{5/4} \end{equation} Figure 5 illustrates the evolution of the disk on these time scales. The top panel shows a blob of material superimposed on the smooth density and temperature structure of a steady disk. In this example, the blob has a mass, $\delta M$, larger than the mass of an annulus, and a thermal energy content, $\delta E$, larger than the local thermal energy of an annulus. Disk rotation broadens the blob into a narrow ring in one or two rotation periods, which produces a narrow spike in the radial distributions of surface density and temperature. The thermal energy of this ring moves inwards and outwards on the thermal time scale. The temperature distribution broadens considerably, but the surface density hardly changes. The mass in the ring finally moves inwards and outwards on the viscous time scale. This evolution produces an extra increase in the temperature due to viscous dissipation. Table 3 compares numerical values for the three disk time scales. These results assume \ifmmode {M_{\star}}\else $M_{\star}$\fi~= 1 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~ and $\alpha = 10^{-1}$. For most interacting binaries, the disk radius is $\sim$ 0.1--10 \ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi; the time scales range from a fraction of a day for $\tau_D$ up to several tens of days for $\tau_V$. The time scales increase dramatically for the larger disks in pre-main sequence stars and active galactic nuclei. The viscous time scale at the edge of a typical solar system is comparable to the disk lifetime. \begin{table}[htb] \begin{center} \caption{Disk time scales} \begin{tabular}{lcccccc} \hline Disk radius & h/r & T (K) & $\tau_D$ & $\tau_T$ & $\tau_V$ \\ \hline 0.1 \ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi & 0.02 & $2 \times 10^4$ & 0.004 d & ~~0.04 d & ~20 d \\ 1.0 & 0.02 & $5 \times 10^3$ & 0.120 d & ~~1.00 d & 400 d \\ \\ 1 AU & 0.04 & 250 & ~~~1 yr & ~~~~~4 yr & $1 \times 10^4$ yr \\ 10 & 0.06 & ~60 & ~~32 yr & ~~~105 yr & $1 \times 10^5$ yr \\ 100 & 0.08 & ~25 & 1000 yr & ~~2500 yr & $2 \times 10^6$ yr \\ \hline \end{tabular} \end{center} \end{table} \subsection{Disk Energy Distributions} Figure 6 shows the broadband spectral energy distribution of a steady accretion disk with a boundary layer. The model assumes that the disk is optically thick and that each annulus radiates as a blackbody (\cite{lyn74}, \cite{bat74}, \cite{tyl81}, \cite{ruc85}). There are two main features in this spectrum: (i) radiation from the boundary layer at short wavelengths, and (ii) disk radiation at longer wavelengths. If the disk has a large ratio of outer radius to inner radius, $R_{out}/R_{in} \gg 1$, the disk spectrum follows \begin{equation} \lambda F_{\lambda} \propto \lambda^{-4/3} ~~ \hspace{20mm} ~~ \lambda_{in} < \lambda < \lambda_{out} \end{equation} \noindent where $\lambda_{in} T_d(R_{in}) \approx \lambda_{out} T_d(R_{out}) \approx$ 0.36 (\cite{lyn74}). The disk spectrum follows the Wien tail of a hot blackbody, $T_d(R_{in})$, at short wavelengths and the Rayleigh-Jeans tail of a cool blackbody, $T_d(R_{out})$, at long wavelengths. Although Figure 6 assumes a blackbody disk, the general features of the model change little if the disk radiates as some type of stellar atmosphere. The most important modifications for pre-main sequence stars involve irradiation from the central star and the infalling envelope, as described by Beckwith in this volume. \epsfxsize=5.in \hskip 12ex \epsffile{fig6.eps} \footnotesize \baselineskip=10pt \vskip -7ex \noindent{Figure 6 - Energy distributions for accretion disks. The left panel shows the energy distribution of a disk (dashed line), a boundary layer (dot-dashed line), and the total energy output (solid line) for \mdot~= $2~\times~10^{-7}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$ onto a solar-type star. The right panel adds the energy distribution (dotted line) for a star with $L_{\star} \approx L_d$. \vskip 3ex \normalsize \baselineskip=12pt Lynden-Bell \& Pringle first applied these concepts to disks in interacting binaries and pre-main sequence stars. They added radiation from the central star to the spectrum, as indicated in the right panel of Figure 6. This model suggests that identifying disk radiation can be difficult when the stellar luminosity exceeds the accretion luminosity. Table 4 compares temperatures and luminosities of pre-main sequence stars and steady disks for two accretions rates, $\mdot = 10^{-6}$ and $10^{-4}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$. At low masses, the contrast between the disk and star is considerable unless the accretion rate is $\mathrel{\copy\simlessbox} 10^{-8}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$. This contrast decreases for more massive stars. At 10 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi, the star is hotter and as luminous as the disk, because massive young stars burn hydrogen before reaching the main sequence. Table 3 shows that accretion cannot be a dominant energy source for main sequence stars; the difficulty of seeing an accretion disk against the background of a young O or B star therefore is not surprising. \begin{table}[htb] \begin{center} \caption{Accretion and stellar luminosities in pre-main sequence stars} \begin{tabular}{lccccccc} \hline & & & & \multicolumn{2}{c}{$10^{-6}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$} & \multicolumn{2}{c}{$10^{-4}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$} \\ \ifmmode {M_{\star}}\else $M_{\star}$\fi (\ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi) & \ifmmode {T_{\star}}\else $T_{\star}$\fi (K) & \ifmmode {L_{\star}}\else $L_{\star}$\fi (\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi) & & $T_{max}$ (K) & $L_{acc}$ (\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi) & $T_{max}$ (K) & $L_{acc}$ (\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi) \\ \hline 0.1 & 3100 & 0.1 & & 1850 & 3 & 5850 & 275 \\ 0.4 & 3600 & 0.5 & & 1825 & 7 & 5800 & 675 \\ 1.0 & 4250 & 7.0 & & 1800 & 12 & 5700 & 1200 \\ 3.0 & 8650 & 80 & & 1700 & 25 & 5350 & 2350 \\ 6.0 & 17000 & 1000 & & 2100 & 50 & 6650 & 5000 \\ 10.0 & 21000 & 5000 & & 1850 & 60 & 5850 & 6000 \\ 20.0 & 29000 & 45000 & & 1600 & 75 & 5000 & 7450 \\ \hline \end{tabular} \end{center} \end{table} \section{Unstable Accretion Disks} Despite the elegant simplicity of the steady-state model, all disks vary their energy output. These fluctuations range from small, 10\%--20\%, amplitude flickering on the local dynamical time scale (\cite{bru94}) to large-scale eruptions, $\sim$ 3--5 mag or more, that can last for several times the local viscous time scale (Figure 1). In well-studied dwarf novae, the outbursts occur in cyclical patterns with quasi-periods of 10 or more days (\cite{can98}, \cite{opp98}). Osaki and Pringle {\it et al.~} first noted the possibility for thermal instabilities in accretion disks (\cite{pri73}, \cite{osa74}). In the standard picture, the structure of a steady disk is derived from balancing heat generated by viscosity with radiative cooling. However, the viscous energy generation and radiative losses are set by local disk parameters; the input \mdot~is an external parameter. Cooling can balance heating only if the local physical quantities can adjust to the input $\mdot$. If the local physical variables cannot adjust, cooling cannot balance heating and a thermal instability is likely. Figure 7 displays a simple illustration of a thermal instability in a single disk annulus. The solid curves indicate loci of stable disks, where cooling precisely balances heating. To the right of these lines, heating exceeds cooling because viscous energy generation -- set by the local surface density $\Sigma$ through $\mdot = \nu \Sigma$ -- exceeds the radiative losses set by the effective temperature, $T_d$. Cooling exceeds heating to the left of the stability lines. To illustrate the evolution of the instability, assume the annulus has ($\Sigma$, $T_d$) at A$^{\prime}$ and an input accretion rate, $\mdot_1$. This input accretion rate exceeds the stable accretion rate at A$^{\prime}$, $\mdot_s = \mdot_{A^\prime}$. The annulus evolves up the solid line on the viscous time scale to find a solution where $\mdot_s = \mdot_1$. Before reaching this point, the annulus arrives at $\mdot_B$, the last stable solution for this opacity source. At B, the annulus needs a larger surface density to accommodate the larger $\mdot$, and a larger $T_d$ to radiate away the extra heat generated by the larger $\mdot$. To do so, the annulus evolves to C; it makes this transition at constant $\Sigma$ because the thermal time scale is shorter than the viscous time scale. Once at C, the annulus radiates more energy than generated by viscous dissipation ($\mdot_s = \mdot_C > \mdot_1$) and evolves down the solid line towards a solution where $\mdot_s = \mdot_1$. The annulus moves towards lower $\Sigma$ on the viscous time scale until it reaches point D, the last stable solution on this branch of the equilibrium curve. At D, the annulus drops to A to try to find the smaller $T_d$ needed to accommodate a smaller $\mdot$. Once at A, the annulus retraces the steps around this `limit-cycle' curve as long as $\mdot = \mdot_1$. \epsfxsize=5.4in \epsffile{fig7.eps} \footnotesize \baselineskip=10pt \vskip -2ex \noindent{Figure 7 - Outline of a disk instability in a single annulus. The disk evolves along the solid lines ABCD for an input $\mdot$ (horizontal arrow). Heavy solid lines correspond to stable equilibrium solutions; a dashed line indicates the locus of unstable equilibria.} \vskip 4ex \normalsize \baselineskip=12pt The evolution of an instability in a complete accretion disk follows the simple illustration. In a real disk, most annuli are close to B when a single annulus makes the transition from B to C. The increased temperature of a single annulus transfers heat to neighboring annuli on the thermal time scale; these annuli then jump to the `high state' and propagate the eruption to their neighbors. Figure 8 shows this evolution for five snapshots in the evolution of a dwarf nova accretion disk. Time increases upwards in the Figure. The bottom panel shows the temperature distribution of the disk at the onset of the eruption. The inner edge of the disk jumps to the high state first; annuli at larger radii progressively follow until all of the disk resides in the high state in the top panel. The disk maintains this state for several days, and then retraces its path back to the low state. \epsfxsize=7.5in \epsffile{fig8.eps} \footnotesize \baselineskip=10pt \noindent{Figure 8 - Evolution of a disk instability in a dwarf nova disk. The bottom panel compares the radial temperature distribution of a time-dependent disk (solid line) with the temperature distribution of a steady disk (dashed line). The eruption begins in the next panel with a sharp increase in the temperature at the inner edge of the disk. This increase creates a `wave' of increased temperature that propagates radially outwards in the disk as time moves forward (up in the figure). In the top panel, the entire disk has reached the hot state; the actual temperature distribution (solid line) is then close to the steady-state temperature distribution (dashed line)} \vskip 4ex \normalsize \baselineskip=12pt One feature of the disk instability picture is that the disk is never in a steady-state. The dashed lines in Figure 8 plot steady-state temperature distributions normalized to the temperature at the inner edge of the disk. The actual temperature gradient is much flatter than the steady-state gradient in the low state (bottom panel). The $\mdot$ through the disk increases radially outwards and is below the input $\mdot$ at each point in the disk. The disk approaches but never reaches the steady-state gradient in the high state (top panel). Actual disk temperature distributions have been derived from eclipse light curves similar to those shown in Figure 1 for CI Cyg. K. Horne \& collaborators (\cite{ho85a}, \cite{ho85b}, \cite{mar88}, \cite{bap98}) have used maximum entropy techniques to recover $T_d(R)$ from multi-wavelength light curves and spectroscopic data assuming (i) the disk is azimuthally symmetric and (ii) the brightness temperature is close to the local blackbody temperature. Their results indicate that quiescent disks are rarely close to steady-state, with temperature distributions usually much flatter than $T_d(R) \propto R^{-3/4}$. Systems in eruption -- dwarf novae and symbiotic stars at maximum light -- more closely resemble, but never achieve, the steady-state temperature distribution with $q$ = 3/4. These results generally agree with the model predictions in Figure 8. \section{Disk Eruptions in Pre-Main Sequence Stars} Most pre-main sequence stars vary in brightness. Irregular brightness variations of $\mathrel{\copy\simlessbox}$ 1 mag are a defining feature of T Tauri stars and many Herbig AeBe stars (\cite{joy45}, \cite{her60}). Recent studies show that many of these variations are due to dark spots rotating with the stellar photosphere (\cite{hrb87}, \cite{bou89}, \cite{bou93}, \cite{hrb94}) or bright spots produced at the base of a magnetic accretion column (\cite{bou89}, \cite{ken94}). Small eruptions of 1--3 mag lasting several years occur in the EXors, a poorly-studied class which includes EX Lup and DR Tau (\cite{her89}). More spectacular 3--6 mag eruptions occur in the FU Orionis variables, also known as FUors (\cite{her77}, \cite{har96}). G. Herbig first associated the eruptions of FUors with pre-main sequence stars (\cite{her66}, \cite{her77}; see also \cite{amb54}, \cite{kho59}). FU Ori, which lies at the apex of a fan-shaped nebula within the dark cloud B35, brightened by a factor of over one hundred in 100--200 days (Figure 9; \cite{wac39}, \cite{wac54}, \cite{her66}, \cite{her77}, \cite{ibr93}, \cite{she95}, \cite{ibr97}). Thirty years later, Welin discovered the 5 mag brightening of V1057 Cyg within an eccentric ring of reflection nebulosity (Figure 9; \cite{wel71}, \cite{her77}, \cite{ken91a}, \cite{kol97}). Herbig later noted the similarity between the optical spectra of these two stars with spectra of V1515 Cyg, a faint variable star embedded in arc-shaped nebulosity (\cite{her60}). He collected archival photographic photometry and identified a slow rise from $m_{pg} \approx$ 15.5 in the late 1940s to $m_{pg} \approx$ 13.5 in the late 1970s (\cite{her77}). This brightness increase continued until 1980, when the star experienced a dramatic decline and slow recovery (\cite{kol83}, \cite{ken91b}). Herbig's demonstration that FU Ori -- and by analogy other FUors -- is a pre-main sequence star is straightforward. FUors are clearly associated with dark clouds, having radial velocities indistinguishable from the cloud velocity. The optical spectrum, including the high lithium abundance, has similarities with spectra of T Tauri stars. The event statistics are plausible for pre-main sequence stars. Finally, the rise in brightness is a real luminosity increase that is {\it not} a nova outburst, the main alternative. \vskip 2ex \epsfxsize=7.0in \epsffile{fig9.eps} \footnotesize \baselineskip=10pt \vskip -6ex \centerline{Figure 9 - Blue light curves for three FU Ori variables.} \vskip 4ex \normalsize \baselineskip=12pt Many new FUors have been discovered since V1515 Cyg. The catalogued number now stands at 11 (\cite{har96}, Table 5). Most have been observed to rise 3--5 mag in optical or near-IR brightness in less than one year, as indicated by the `Y' in the outburst (OB) column of Table 5. V1515 Cyg is the only known example to require a decade to rise to visual maximum, but the historical light curves for some systems are poorly documented. A few objects have been called FUors based on common spectroscopic properties, which are described more completely below. Most recent FUors are more intimately associated with the densest dark clouds than the first members of the class, which suggests that many eruptions might have been missed. There are also 3 candidate FUors listed as a second group of objects in Table 5. These objects have some properties in common with FUors, but more data are needed to see if they have the other characteristics as well (\cite{str93}, \cite{san98}). A few other objects have one or two properties of known FUors, such as an outburst or unusual light curve (e.g., \cite{eis91}, \cite{hod96}, \cite{alv97}, (\cite{she97}, \cite{yun97}). These are more probably EXors or Ae/Be stars than FUors. \subsection{Basic Properties of FU Orionis Objects} FUors share a distinctive set of morphological, photometric, and spectroscopic characteristics (Table 5). Three are known binary stars, L1551 IRS5, Z CMa, and RNO 1B/1C (\cite{kor91}, \cite{ken93b}, \cite{bar94}, \cite{thi95}, \cite{rod98}). Most have delicate fan-shaped or comma-shaped reflection nebulae (RN) on optical and near-IR images (\cite{goo87}, \cite{rei91}, \cite{nak95}, \cite{luc96}). Optical jets and HH objects are common (\cite{rei91}, \cite{dav94}, \cite{har96}). Most appear associated with large-scale molecular outflows (\cite{yam92}, \cite{fri93}, \cite{eva94}, \cite{mcm95}, \cite{yan95}, \cite{fri98}, \cite{lop98}, and references therein). These features -- together with the broad, blue-shifted Na~I and H~I absorption features described below -- demonstrate that FUors drive powerful winds that interact with the surrounding medium (\cite{bas85}, \cite{cro87}, \cite{whi93}). \begin{table}[htb] \begin{center} \caption{Selected properties of FU Orionis objects} \begin{tabular}{lcccccccc} \hline Object & OB & Opt ST & IR ST & RN & HH/Jet & Wind & Outflow & Radio \\ \hline L1551 IRS 5 & ? & K & M & Y & Y & ? & Y & Y \\ FU Ori & Y & F--G & M & Y & N & Y & N & N \\ Z CMa & ? & F--G & M & Y & Y & Y & Y & Y \\ BBW 76 & @ & G & M & Y & ? & Y & N & ? \\ V346 Nor & Y & ? & C & Y & Y & ? & Y & ? \\ Par 21 & ? & F & M & Y & Y & Y & ? & ? \\ V1515 Cyg & Y & G & M & Y & N & Y & Y & N \\ V1057 Cyg & Y & G & M & Y & N & Y & Y & Y \\ V1735 Cyg & Y & ? & M & Y & N & ? & Y & Y \\ RNO 1B/1C & Y & G & M & Y & N & ? & Y & ? \\ \\ IC 430 & ? & G & ? & Y & Y & Y & N & Y \\ PP 13S & Y & ? & M & ? & ? & ? & Y & ? \\ Re 50 & ? & G & ? & Y & Y & Y & Y & Y \\ \hline \end{tabular} \end{center} \end{table} FUors also have very unusual spectroscopic characteristics. The optically visible sources have F--G, and sometimes K-type, giant or supergiant spectra (\cite{her77}; Table 5, see also \cite{har88}, \cite{re97a}, \cite{re97d})). The optical reflection nebulae of several embedded FUors also show G-type absorption features (\cite{sto88}, \cite{sta92}). All FUors but Z CMa and V346 Nor have very deep CO absorption bands on near-IR spectra at 1.6 $\mu$m and 2.3 $\mu$m (\cite{mou78}, \cite{eli78}, \cite{car87}, \cite{sat92}, \cite{ken93a}, \cite{bis97}). These features resemble the CO absorption bands observed in red giants and are much stronger than those observed in any other pre-main sequence star (\cite{har87a}, \cite{har87b}, \cite{gre97}). CO absorption in Z CMa is weakened by dust emission from an embedded companion; V346 Nor has weak CO emission instead of CO absorption (\cite{kor91}, \cite{teo97}; \cite{re97d}). Many FUors also display strong water absorption features, which strengthens the evidence for a $\sim$ 2000 K photosphere (\cite{cal91}, \cite{sat92}, \cite{cal93}). All FUors show large {\it excesses} of radiation over normal G supergiants at both ultraviolet and infrared wavelengths. The near-IR excess is clearly photospheric in origin, because the CO and H$_2$O absorption features are so intense. The UV excesses in Z CMa and FU Ori appear associated with an A- or F-type photosphere that is hotter than the G-type photosphere observed at longer wavelengths (\cite{ke89b}). In addition to significant far-IR and submm emission (\cite{wei89}, \cite{wei91}, \cite{ken91a}), many FUors are strong radio continuum sources at cm wavelengths (\cite{coh82}, \cite{rod90}, \cite{rod92}, \cite{loo97}, \cite{rod98}). This emission is not photospheric in some FUors and may be produced in the outflow or the jet. Herbig first noted broad absorption lines on optical spectra of several FUors (\cite{her77}). Hartmann \& Kenyon later confirmed large rotational velocities of $v$ sin $i$ $\approx$ 15--60 \ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi~for V1057 Cyg, V1515 Cyg, and FU Ori (\cite{har85}). This property is now characteristic of the class; all FUors have broad optical or infrared absorption lines or both (\cite{har88}, \cite{sta91}, \cite{sta92}, \cite{ken93a}). The rotational velocity further depends on wavelength. The near-IR CO lines in FU Ori and V1057 Cyg have significantly smaller rotational velocities than the optical lines (\cite{har87a}, \cite{har87b}). The optical rotational velocity smoothly increases with decreasing wavelength in Z CMa and V1057 Cyg (\cite{wel91}, \cite{wel92}). In FU Ori itself, a powerful wind masks weak absorption lines and makes it difficult to detect any variation of rotational velocity with wavelength should one exist. Many FUors display {\it doubled} absorption lines on optical and near-IR spectra (\cite{har85}, \cite{har87a}, \cite{har87b}, \cite{har88}, \cite{sta91}, \cite{sta92}). The two absorption components in V1057 Cyg are separated by 30--40 \ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi; these features have much larger separations in Z CMa and FU Ori. The long-term stability of the doubled absorption lines indicates that the lines are not produced by two stellar components in a binary system (\cite{ken88}). Most FUors also show strong evidence for mass loss, in addition to the HH objects and CO outflows described above. Practically every FUor displays deep, blueshifted absorption components on H$\alpha$ and Na I D (\cite{bas85}, \cite{cro87}, \cite{eis90}, \cite{wel92}, \cite{har95}); weak blueshifted CO absorption might be present in FU Ori (\cite{har87b}). The optical line profiles can also change on month to year time scales (\cite{bas85}, \cite{cro87}, \cite{wel92}). The near-IR line profiles may also change (\cite{bis97}). Both V1057 Cyg and V1515 Cyg have very pronounced line profile changes and dips in their optical light curves. These fluctuations suggest dust formation in a variable wind, but the data are not very extensive (\cite{ken91b}). Finally, FUor eruptions must be repetitive (\cite{her77}, \cite{har85}). The event statistics of known FUors suggest that a young star must undergo 10--20 FUor eruptions before reaching the main sequence. This estimate may be a lower limit, because some outbursts have certainly been missed. Reipurth's proposal that FUor eruptions produce HH objects suggests a recurrence time scale of $\sim$ 1000 yr, based on the identification of multiple bowshocks (ejection events) with dynamical separations of 500--2000 yr (\cite{rei85}, \cite{rei89}, \cite{hrt90}, \cite{rei92}, \cite{bac94}). The discovery of giant HH flows with dynamical time scales of $10^4$--$10^5$ yr (\cite{re97c}) allows 10--100 eruptions if eruptions actually power the outflow (see below). \subsection{FU Orionis Objects as Accretion Disks} The observations described above place severe constraints on possible FUor outburst mechanisms. With a luminosity of a few hundred \ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi~(Table 6), a typical FUor emits $10^{45}$ to $10^{46}$ erg during the course of a 10--100 yr eruption. Outbursts also recur every $10^3$--$10^5$ yr. This process produces an object resembling a rapidly rotating F-G supergiant in the optical and a more slowly rotating M giant in the infrared. It must be unique to young stars, because we have not observed a FUor event in an older star system. Accretion is the most plausible energy source for FUor eruptions (\cite{har85}, \cite{lin85}). By analogy with dwarf novae, FUor eruptions are the high state of a thermal instability in a disk surrounding a pre-main sequence star. This instability occurs if the accretion rate from the molecular cloud core into the disk lies between the stable accretion rates in the low state and the high state (Figure 7). The system cycles between these two states as long as the surrounding cloud can replenish the disk between outbursts. \begin{table}[htb] \begin{center} \caption{Disk Properties of FU Orionis Objects} \begin{tabular}{lccc} \hline Object & $L_{bol}$ (\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi) & \ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot ($10^{-5}~\ifmmode {\rm M_{\odot}^2\,yr^{-1}}\else $\rm M_{\odot}^2\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$) & \ifmmode {R_{\star}}\else $R_{\star}$\fi (\ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi) \\ \hline L1551 IRS 5 & 30 & 0.1 & 0.4 \\ FU Ori & 220 & 1.7 & 1.2 \\ Z CMa & 400 & 4.0 & 1.6 \\ BBW 76 & 200 & 1.4 & 1.1 \\ V346 Nor & 160 & 1.0 & 1.0 \\ Par 21 & 200 & 1.4 & 1.1 \\ V1515 Cyg & 135 & 0.8 & 0.9 \\ V1057 Cyg & 400 & 4.0 & 1.6 \\ V1735 Cyg & 200 & 1.3 & 1.1 \\ RNO 1B/1C & 750 & 3.5 & 1.5 \\ \hline \end{tabular} \end{center} \end{table} Table 6 lists the disk properties needed to power observed FUor luminosities from accretion. To make these estimates, I assume a steady disk with $L_{disk} = 157~\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi (\ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot / \ifmmode {R_{\star}}\else $R_{\star}$\fi) $ and $T_{max} \approx$ 6500 K $(\ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot / \ifmmode {R_{\star}}\else $R_{\star}$\fi^3)^{1/4} $. Combining these and setting $T_{max}$ = 6500 K for an F-G supergiant star yields \begin{equation} \ifmmode {R_{\star}}\else $R_{\star}$\fi \approx \left ( \frac{L_{disk}}{157~\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi} \right )^{1/2} \end{equation} \noindent and \begin{equation} \ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot \approx \left ( \frac{L_{disk}}{157~\ifmmode {\rm L_{\odot}}\else $\rm L_{\odot}$\fi} \right )^{3/2} \end{equation} \noindent These expressions produce the results in Table 6 for known FUors, assuming a typical inclination angle, cos $i$ = 1/2. The median inner radius for the FUor sample is \ifmmode {R_{\star}}\else $R_{\star}$\fi $\approx$ 1.1--1.2 \ifmmode {\rm R_{\odot}}\else $\rm R_{\odot}$\fi; the median mass accretion rate is $\ifmmode {M_{\star}}\else $M_{\star}$\fi \mdot \sim$ 1--2 $\times~10^{-5}~\ifmmode {\rm M_{\odot}^2\,yr^{-1}}\else $\rm M_{\odot}^2\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$. For a mass-radius relation for pre-main sequence stars with ages of $10^5$ yr, $\ifmmode {M_{\star}}\else $M_{\star}$\fi \approx$ 0.2--0.3 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~and $\mdot \approx$ 0.5--1 $\times~10^{-4}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$. This very simple analysis indicates that a young star must accrete material at $\sim 10^{-4}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$ during an eruption to power the observed luminosity. For an adopted outburst duration of 100 yr, this accretion rate results in a total accreted mass of $\sim$ 0.01 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~per eruption. The total mass accreted during a FUor eruption must be replenished during a low state lasting $\sim$ $10^3$--$10^5$ yr, which results in an infall rate of $\sim$ $10^{-5}$--$10^{-7}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$. This range is close to the infall rates envisioned -- and in some cases observed -- for typical cloud cores (\cite{ada87}, \cite{oha96}, \cite{mar97}). \subsection{Observational Tests of Disk Models} Steady accretion disk models successfully explain many observations of FUors (\cite{har85}, \cite{ken88}, \cite{har96}). Theoretical models of disk instabilities produce eruptions that generally resemble FUor events, although comparisons with observations are still in their early stages (\cite{lin85}, \cite{cla89}, \cite{cla90}, \cite{kaw93}, \cite{bel94}, \cite{bel95}, \cite{cla96}, \cite{oku97}, \cite{tur97}). The next few paragraphs summarize these results; Herbig \& Petrov present a different interpretation (\cite{her92}, \cite{pet98}; see also \cite{lar80}). Figure 10 shows light curves for V1057 Cyg, one of the best-studied FUors. These light curves closely resemble those of other accreting systems and provide strong support for an accretion model (\cite{ken88}, \cite{cla90}, \cite{ken91a}, \cite{bel95}, \cite{cla96}, \cite{lin96}, \cite{tur97}). The amplitude of the decline is largest in the UV and decreases monotonically from 0.3 $\mu$m to 5 $\mu$m. This behavior is characteristic of accreting systems, which evolve towards cooler temperatures as the luminosity declines (\cite{ken84}, \cite{bel95}, \cite{lin96}, \cite{can98}). The spectral type variation also agrees with disk model predictions. If we assign stellar effective temperatures to the optical spectrum at maximum (A5 star) and at the current epoch (G5 star), the UBV decline indicates a source with a roughly constant radius. This requirement is easy to achieve with an accretion disk if the radius of the central star remains constant. Most stars cannot evolve at constant radius if their effective temperatures change. \epsfxsize=7.0in \epsffile{fig10.eps} \vskip -12ex \footnotesize \baselineskip=10pt \noindent Figure 10 - Light curves for V1057 Cyg. The symbols alternate between filled circles (at B, R, and K) and plus signs (V and J). The system varied erratically at B = 14--17 prior to outburst and then rose 5--6 mag in less than one year. The visual observations indicate a similar rise time, although few pre-outburst observations are available. No pre-outburst near-IR data exist; K-band observations began shortly after the outburst was detected. The amplitude of the decline clearly depends on wavelength, with larger amplitudes at shorter wavelengths (see \cite{ken91a}, \cite{she95}, \cite{kol97}). \vskip 3ex \normalsize \baselineskip=12pt The light curves of other FUors also generally agree with disk model predictions. The rise times of 200--400 days in FU Ori and V1057 Cyg are comparable to the thermal time scale, as expected for an eruption that begins at the inner edge of the disk (Table 3; \cite{bel95}). The several decade rise in V1515 Cyg is too long for such an ``inside-out'' instability, but could be caused by an outburst that begins in the outer part of the disk . These ``outside-in'' eruptions can result from perturbations of the disk by a companion star or planet, in addition to the disk instability mechanism described above (\cite{cla90}, \cite{bon92}, \cite{bel95}, \cite{cla96}). In all systems, the decay times are comparable to the viscous time scale (Table 3). \epsfxsize=6.0in \epsffile{fig11.eps} \footnotesize \baselineskip=10pt \vskip -2ex \noindent Figure 11 -- Spectral energy distribution for FU Ori. The observations (filled circles) indicate a clear excess over a G-type star (dashed line). A disk model that produces a G-type optical spectrum accounts for most of the data (solid line). The modest far-IR excess over the disk spectrum at $\sim$ 100 \ifmmode {\rm \mu {\rm m}}\else $\rm \mu {\rm m}$\fi~is probably produced by the surrounding nebula responsible for the fan-shaped reflection nebula. \vskip 3ex \normalsize \baselineskip=12pt Disk models also successfully explain FUor SEDs (\cite{har85}, \cite{ken88}, \cite{ken91a}). Figure 11 shows a dereddened SED for FU Ori. The deep UV and near-IR absorption features indicate that the UV and IR excesses over a G-type supergiant spectrum are photospheric (\cite{cal91}). If disks radiate like stars at the effective temperatures appropriate for the observed absorption features, the SEDs for FU Ori and other FUors require that the surface area of the emitting region increases with increasing wavelength. In particular, the surface area of the near-IR source must be 10--20 times larger than the surface area of the optical source; the 300 K material responsible for the 10 $\mu$m excess must have roughly 100 times the emitting area of the optical continuum region. A disk in which the temperature decreases radially outward (\S2) naturally explains this observation. Disk models account for the 0.3--10 $\mu$m SEDs of V1057 Cyg, V1515 Cyg, and FU Ori quite well (Figure 11; see also \cite{har85}, \cite{ada87}, \cite{ken88}, \cite{ken91a}, \cite{bel95}, \cite{tur97}). Several other FUors with modest reddening have similar SEDs (\cite{ken91a}). Recent interferometric observations have resolved FU Ori at near-IR wavelengths. The angular size at 2.2 \ifmmode {\rm \mu {\rm m}}\else $\rm \mu {\rm m}$\fi~agrees with predictions of the accretion disk model (\cite{mal98}). The disk model {\it fails} to explain 10--100 $\mu$m data of many FUors, although the model fits both FU Ori and BBW 76 from 1--100 $\mu$m (\cite{ken91a}). However, the decline of the 10--20 $\mu$m light of V1057 Cyg follows the optical light curve very closely (\cite{ken91a}). This behavior suggests that the mid-IR radiation is optical light absorbed and reradiated by a surrounding envelope. An infall rate of 1--5 $\times~10^{-6}$ \ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi~produces an optically thick envelope that can reprocess optical light from the inner disk and account for the 10--100 $\mu$m SED (\cite{ken91a}, \cite{bel95}). This rate is sufficient to replenish the disk in 1000 yr, which allows recurrent FUor eruptions in this system. The observational evidence for infall at comparable rates in two FUors, Z CMa (\cite{lil97}) and L1551 IRS5 (\cite{oha96}, \cite{mar97}), lends support to this interpretation. Finally, an accretion disk also naturally explains the optical and IR line profiles observed in some FUors (\cite{har87a}, \cite{har87b}, \cite{ken88}, \cite{har88}, \cite{wel91}, \cite{wel92}). The gradual decrease in the rotational velocity with increasing wavelength occurs because the longer wavelength emission is produced in more slowly-orbiting material at larger disk radii than the more rapidly moving inner disk material responsible for the short wavelength emission. At a given disk temperature, a larger fraction of the disk surface rotates at high line-of-sight velocities and the lines appear doubled (\cite{ken88}, \cite{cal93}, \cite{pop96}). \subsection{The Importance of FU Ori Eruptions} Despite their relative rarity, FUors are an important part of pre-main sequence stellar evolution. FUor eruptions add a significant fraction of a stellar mass to the central star. FUor eruptions also eject a large amount of mass into the surrounding cloud. FUor eruptions may be frequent enough to provide nearly all of the mass of a typical low mass star. If this hypothesis is correct, FUor eruptions can power most molecular outflows and may be a dominant source of cloud heating in molecular cloud cores. This view is controversial, but can be tested with observations as outlined below. Although the statistics are still crude, young stars clearly accrete much of a stellar mass in FUor events. A typical eruption adds $\sim$ 0.01 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~to the central star; a rate of 10--20 eruptions per pre-main sequence star leads to a total accreted mass of $\sim$ 0.1--0.2 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi. These eruptions eject considerable amounts of mass and momentum into the surrounding cloud. For example, FU Ori itself has lost material with a momentum of $\sim$ 0.3 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~\ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi~since its eruption began. With 10--20 such eruptions, FU Ori could supply a small fraction of the momentum of a typical molecular outflow, $(Mv)_o~\sim$ 1--100 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~\ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi~(\cite{fuk89}). Two conditions must be satisfied for a young star to accrete {\it all} of its mass in FUor events: 1. The disk must undergo regular thermal instabilities and cycle between the low and high states on a reasonable time scale. 2. FUors must be young, $\mathrel{\copy\simlessbox}$ a few $\times~10^5$ yr, to allow the surrounding cloud to fuel the recurring instabilities. As a group, FUors clearly are young objects. FUors have many more properties in common with young class I protostars than older optically visible T Tauri stars. FUors have distinctive reflection nebulae and large far-IR excesses; most are associated with jets, HH objects, and molecular outflows. These properties are common in class I sources and rare in T Tauri stars (\cite{rei85}, \cite{rei90}, \cite{gom97}). The typical age of a FUor is therefore $\sim$ a few $\times~10^5$ yr (\cite{ken95}, \cite{wei91}, \cite{ken91a}). The conditions needed for the disk to cycle between the high and low states are probably satisfied in many pre-main sequence stars. For a typical infall rate of 1--10 $\times$ $10^{-6}~\ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi$, theoretical models indicate that the disk spends most of its time in a low state where the mass accretion rate through the disk is much lower than the infall rate (\cite{bel94}, \cite{bel95}, \cite{lin96}). The disk requires only $\sim$ $10^3$--$10^4$ yr to accumulate the $\sim$ 0.01 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~needed to power a FUor eruption. The disk will continue to cycle between the low and high states as long as the cloud supplies mass to the disk. If young stars accrete most of their mass in FUor events, it follows that FUors can power molecular outflows. In most outflow models, the mass ejected by the central star is $\sim$ 10\%--30\% of the accreted mass. For a wind velocity of $\sim$ 200~\ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi, the momentum in the wind is \begin{equation} (Mv)_w ~ \sim ~ 20 ~ (\ifmmode {M_{\star}}\else $M_{\star}$\fi/\ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi) ~ \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~\ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi \end{equation} \noindent for a typical wind velocity of 200 \ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi. The observed range of outflow momenta -- $(Mv)_o$ = 1--100 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~\ifmmode {\rm km~s^{-1}}\else ${\rm km~s^{-1}}$\fi~(\cite{fuk89}) -- requires stellar masses of $\ifmmode {M_{\star}}\else $M_{\star}$\fi ~ \sim$ 0.1--5 \ifmmode {\rm M_{\odot}}\else $\rm M_{\odot}$\fi~if the flows conserve momentum. The winds of FUors can power this outflow {\bf if} a young star accretes most of its mass in FUor events. These simple estimates establish that FUors are important events if they recur on time scales of $\sim 10^3$ yr during the protostellar phase of pre-main sequence stellar evolution. In this picture, the fraction of time in the FUor state is simply the ratio of the infall rate to the FUor accretion rate: $f \sim \mdot_i/\mdot_{FU}$. FUor disk models require $\mdot_{FU} \sim 10^{-4}$ \ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi, so $f \sim 0.05$ for a typical infall rate of $\mdot_i \sim 5 \times 10^{-6}$ \ifmmode {\rm M_{\odot}\,yr^{-1}}\else $\rm M_{\odot}\,yr^{-1}$\fi}\def\mdot{\ifmmode {\dot{M}}\else $\dot{M}$\fi. This prediction currently agrees with the statistics of FUors among protostars: 1 out of $\sim$ 20 protostars in the Taurus dark cloud contains a FUor (L1551 IRS5; \cite{ken95}) and 5 out of 17 protostars with HH objects contains a central star that resembles known FUors spectroscopically (\cite{re97b}; see also \cite{han95}). Sensitive photometric and spectroscopic surveys of nearby molecular clouds can improve these statistics considerably. If the FUor frequency among protostars turns out to be more than a few per cent, then FUors may represent the main accretion phase of early stellar evolution. I thank Janet Mattei for the SS Cyg data and John Cannizzo for the disk instability models.
\section{INTRODUCTION} Interaction between an accretion disk and a magnetized central object lies at the heart of the physics of a variety of astrophysical systems, including accreting neutron stars, white dwarfs and pre-main-sequence stars (e.g., Frank et al.~1992; Hartmann 1998). The basic picture of disk-magnetosphere interaction was first outlined by Pringle \& Rees (1972) following the discovery of accretion-powered X-ray pulsars. These are rotating, highly magnetized ($B\sim 10^{12}$~G) neutron stars that accrete material from a companion star, either directly from a stellar wind, or in the form of an accretion disk. The strong magnetic field disrupts the accretion flow at the magnetospheric boundary (typically at a few hundreds neutron star radii), and channels the plasma onto the polar caps of the neutron star. The magnetosphere boundary is located where the magnetic and plasma stresses balance, \begin{equation} r_m=\eta \left({\mu^4\over GM\dot M^2}\right)^{1/7}, \label{alfven}\end{equation} where $M$ and $\mu$ are the mass and magnetic moment of the central object, $\dot M$ is the mass accretion rate, $\eta$ is a dimensionless constant of order unity (For an aligned dipole, estimate of $\eta$ ranges from $0.5$ to $1$; e.g., Pringle \& Rees 1972; Lamb, Pethick \& Pines 1973; Ghosh \& Lamb 1979a,b,~1992; Aly 1980; Arons 1993; Wang 1995). In low-mass X-ray binaries containing weakly magnetized ($B\sim 10^8$~G) neutron stars, the magnetosphere lies close to the stellar surface; the neutron stars are spun up and eventually become millisecond pulsars (Alpar et al.~1982; see Phinney \& Kulkarni 1994 or Bhattacharya 1995 for a review). Similar magnetosphere-disk interaction also occurs in T Tauri stars (e.g., Hartmann 1998), where the stellar magnetic field ($B\sim 10^3$~G at the surface) strongly affects the accretion flow, and provides a key ingredient to explain the anomalous slow rotation of these objects (e.g., K\"onigl 1991; Cameron \& Campbell 1993; Shu et al.~1994). Finally, models of magnetized accretion disks have been applied to explain the observed UV and X-ray emission from some intermediate polars (the DQ Her subclass of CVs) (e.g., Yi \& Kenyon 1997). A large number of theoretical papers have been written on the subject of the interaction between accretion disk and a magnetized star (e.g., Pringle \& Rees 1972; Ghosh \& Lamb 1979a,b; Aly 1980; Lipunov \& Shakura 1980; Anzer \& B\"orner 1980,1983; Arons 1987,1993; Wang 1987,1995; Aly \& Kuijpers 1990; Spruit \& Taam 1990,1993; Shu et al.~1994; van Ballegooijen 1994; Lovelace et al.~1995,1998; Li, Wickramasinge \& R\"udiger 1996; Campbell 1997), and numerical study of this problem is still in its infancy (e.g., Stone \& Norman 1994; Hayashi et al.~1996; Goodson et al.~1997; Miller \& Stone 1997; see also Toropin et al.~1999). Beyond the simple notion of magnetosphere (with the Alf\'ven radius estimated by eq.~[\ref{alfven}]), there is little consensus on the range and strength of magnetic interaction outside the magnetosphere boundary, on the efficiency of magnetic field dissipation in/outside the disk, and on how or where the plasma attaches to the field lines. Outstanding issues include whether the disk excludes the stellar magnetic field by diamagnetic currents or the field can penetrate and thread a large fraction of the disk, whether the threaded field remains closed (connecting the star and the disk) or becomes open by differential shearing, and whether/how magnetically driven wind is launched from the disk or the magnetosphere/corotation boundary. This paper introduces some new physical effects associated with disk accretion onto rotating magnetized stars. Most previous papers have, for simplicity, assumed that the stellar magnetic dipole is aligned with the spin axis. Essentially all previous authors, with the exception of Lipunov \& Shakura (1980) (see \S 2.1), have assumed that the spin is aligned with the disk angular momentum --- There is a good reason for this, as it is natural to think that the star attains its spin angular momentum from the accretion disk. However, when these assumptions are abandoned, new physical effects are revealed: We show that under quite general conditions, the stellar magnetic field can induce disk warping and make the disk precess around the spin axis. As the accreting plasma approaches the magnetosphere, there is a tendency for the disk plane to be driven away from the equatorial plane of the star toward being aligned with the spin axis (\S 2 and \S 4). The origin of the {\it precessional torque} and {\it magnetic warping instability} lies in the interaction between the surface current on the disk and the horizontal magnetic field (parallel to the disk) produced by the stellar dipole. The electric current on the disk is an inevitable consequence of magnetic-field -- disk interaction, although its actual form is uncertain and depends on whether the disk is diamagnetic or has a vertical field threading it, thus on the dissipative processes in the disk and the magnetosphere (\S 2). The magnetic force on the disk also give rise to vertical and epicyclic resonances which may drive disk perturbations (\S 3). We show that the magnetically driven warping and precession of accretion disks can potentially explain a number of outstanding puzzles related to disk accretion onto magnetized stars (\S 6). This paper is organized as follows. In \S 2 we calculate the precessional and warping torques on the disk (or a ring of disk) due to the stellar magnetic field for several models based on different assumptions about magnetic field -- disk interaction. The torques are different for different models, but they are always present and are of the same order of magnitude. The key physics of the torques are summarized at the beginning of \S 2. Resonances due to magnetic forces on the disk are discussed in \S 3. In \S 4 we study the dynamics of a warped disk under the influence of the magnetic torques derived in \S 2. The criterion for the magnetically driven warping instability (including the effect of viscosity) is derived, and a magnetic ``Bardeen-Petterson'' effect is discussed. Section 5 addresses the question of how the spin of the central star is affected by the warped, precessing disk. In \S 6 we discuss/speculate several astrophysical applications of our theory, including the spin evolution of accretion-powered X-ray pulsars, quasi-periodic oscillations in low-mass X-ray binaries, long-term periodic cycles in several X-ray binaries, and the variability and rotation of T Tauri stars. We conclude in \S 7 by discussing possible future studies along the line initiated in this paper. Because of the intrinsic uncertainties associated with the nature of magnetic field -- disk interaction, in the main text we shall focus on generic features and rely on parametrized models. In Appendix A we discuss a global magnetized disk model. In Appendix B we consider the issue of calculating the magnetosphere radius for nontrivial magnetic field and spin geometry. The dynamics of an extreme form of disk (consisting of diamagnetic blobs) is examined in Appendix C. \section{MAGNETIC PRECESSIONAL AND WARPING TORQUES} In this section we calculate the magnetic torque on the disk (or a ring of disk) surrounding a rotating magnetic dipole. Because of the uncertainties on how the magnetic field behaves in the presence of an accretion disk (see \S 1), we shall consider several different models, some representing extreme situations. Our most ``realistic'' model is given in \S 2.3. The basic set-up used in our calculation is shown in Fig.~1. The disk angular momentum axis ${\hat l}$ is inclined by an angle $\beta$ with respect to the spin axis ${\hat\omega}$. The stellar dipole momentum ${\hat\mu}$ rotates around the spin axis ${\hat\omega}$. The angle of obliquity between the magnetic moment and the rotation axis is $\theta$. The key physics responsible for the magnetic torques on the disk is as follows: (i) If the disk is diamagnetic (perfectly conducting) so that the vertical stellar field (perpendicular to the disk) cannot penetrate, an azimuthal screening surface current will be induced in the disk. This current interacts with the radial magnetic field from the star, and a vertical magnetic force results. While the mean force (averaging over the azimuthal) is zero, the uneven distribution of the force induces a net torque acting on the ring, making it precess around the spin axis. (For a nonrotating star, the disk will precess around the magnetic axis.) (\S 2.1). (ii) If there is a vertical field threading the disk (this $B_z$ either comes from the star or is carried intrinsically by the disk and detached from the star), it will be twisted by the disk rotation to produce a discontinuous azimuthal field $\Delta B_\phi$ across the disk surface, and a radial surface current results. The interaction between this current and the stellar $B_\phi$ (which is not affected by the disk) gives rise to a vertical force, and the resulting torque tends to misalign the angular momentum of the ring with the spin axis (although in some extreme situations, alignment torque may also result). (For a nonrotating star, the disk normal will be driven away from the stellar magnetic axis.) (\S 2.2). In general, we expect both types of torques to exist on the disk (\S 2.3 and \S 2.4). \subsection{Diamagnetic Disk} Here we consider the extreme situation where the disk is a perfect conductor and has no large-scale magnetic field of its own. The inner radius of the disk is located at $r=r_m$, the magnetospheric radius. The magnetic field produced by the stellar dipole cannot penetrate the disk, and a diamagnetic surface current is induced. Aly (1980) has found the exact analytic solution to this model problem (see also Kundt \& Robnik 1980; Riffert 1980). The magnetic field at a point $(r,\phi,z=0)$ (cylindrical coordinates) on the disk surface ($z=0,~r>r_m$) is given by \begin{eqnarray} B_r&=&{2\mu\over r^3}\sin\chi\cos(\phi-\phi_\mu)\mp{4\mu\over\pi r^3D} \cos\chi,\label{br1}\\ B_\phi&=&{\mu\over r^3}\sin\chi\sin(\phi-\phi_\mu),\label{bphi1}\\ B_z&=&0, \end{eqnarray} where $\chi$ and $\phi_\mu$ are defined in Fig.~1 (they are both varying in time). In eq.~(\ref{br1}), the upper (lower) sign applies to the upper (lower) disk surface, and the factor $D$ is \begin{equation} D={\rm max}\left(\sqrt{{r^2/r_m^2}-1}, \sqrt{{2H/r_m}}\right), \label{Dfunc}\end{equation} (where $H\ll r_m$ is the half-thickness of the disk). The discontinuity in $B_r$ implies a surface current \begin{equation} {\bf K}=-{2c\mu\over \pi^2r^3D}\cos\chi\,\hat\phi \end{equation} ($\hat\phi$ is the unit vector along the $\phi$ direction; similar notation will be used throughout the paper). Note that this surface current is induced to cancel the $z$-component of the stellar field, $B_z^{(0)}=-\mu\cos\chi/r^3$; the $r,\phi$-components of the stellar field [the first terms of eqs.~(\ref{br1}) and (\ref{bphi1})] do not induce any net surface current. (In fact, they induce currents on the upper and lower surfaces of the disk, but these currents have opposite directions. Such currents can lead to the ``squeezing'' of the disk --- changing the thickness of the disk, but not the ``lifting''.) The magnetic force per unit area on the disk results from the interaction between ${\bf K}$ and $B_r^{(0)}$ (the first term in eq.~[\ref{br1}]): \begin{equation} {\bf F}={2\mu^2\over\pi^2r^6D}\sin 2\chi\cos(\phi-\phi_\mu)\,\hat z. \label{force1}\end{equation} The existence of this vertical magnetic force has already been noted by Aly (1980), and it is simply the difference in the magnetic pressure, $B^2/(8\pi)$, between the lower and upper surfaces. The magnetic torque per unit area is \begin{equation} {\bf N}={\bf r}\times {\bf F}=-{2\mu^2\over\pi^2r^5D}\sin 2\chi\cos(\phi-\phi_\mu)\hat\phi. \end{equation} Clearly, averaging the force over the azimuthal angle, $\langle{\bf F}\rangle_\phi =(1/2\pi)\int_0^{2\pi}d\phi\, {\bf F}$, gives zero, but the azimuthally averaged torque is nonzero, and given by \begin{equation} \langle{\bf N}\rangle_\phi={\mu^2\over\pi^2r^5D}\sin 2\chi(\sin\phi_\mu{\hat x} -\cos\phi_\mu{\hat y}) ={2\mu^2\over\pi^2r^5D}\cos\chi\,({\hat\mu}\times{\hat l}), \end{equation} (${\hat\mu}$ and ${\hat l}$ are the unit vectors along the dipole moment and the disk angular momentum, respectively). For nonrotating star, this result implies that the magnetic torque tends to make the disk precess around the magnetic axis ${\hat\mu}$. For a rotating star --- as long as the rotation period is much shorter than the precession period, we need to average over the spin period. The following identities will be needed: \begin{eqnarray} \cos\chi &=&\cos\beta\cos\theta-\sin\beta\sin\theta\sin\omega t,\label{id1}\\ \sin\chi\cos\phi_\mu &=&\sin\theta\cos\omega t,\label{id2}\\ \sin\chi\sin\phi_\mu &=& \sin\beta\cos\theta+\cos\beta\sin\theta\sin\omega t, \label{id3}\end{eqnarray} where an appropriate phase for the stellar rotation has been adopted. We find that, after averaging over the spin period, the torque per unit area $\langle\langle{\bf N}\rangle\rangle\equiv \langle\langle{\bf N}\rangle_\phi\rangle_\omega\equiv (1/P_s)\int_0^{P_s}dt\,\langle{\bf N}\rangle_\phi$ (where $P_s=2\pi/\omega$ is the spin period) is given by \begin{equation} \langle\langle{\bf N}\rangle\rangle ={\mu^2\over \pi^2r^5D}\cos\beta \left(3\cos^2\theta-1\right)\,({\hat\omega}\times{\hat l}). \label{torque1}\end{equation} The magnetic torque on a ring of radius $r$ and width $dr$ is simply $d{\bf T}=2\pi r\langle\langle{\bf N}\rangle\rangle dr$. The angular momentum of the ring is $(2\pi r\Sigma dr)(r^2\Omega)$ (where $\Sigma$ is the surface density of the disk, and $\Omega$ is the orbital angular frequency). Thus the effect of the magnetic torque is to make the ring precess around the spin axis ${\hat\omega}$ at an angular frequency \begin{equation} {\bf \Omega}_{\rm prec}={\mu^2\over \pi^2r^7\Omega\Sigma D}\cos\beta\, \left(3\cos^2\theta-1\right)\,{\hat\omega}. \end{equation} Note that the sign of $\cos\beta \left(3\cos^2\theta-1\right)$ determines whether ${\bf \Omega}_{\rm prec}$ is along ${\hat\omega}$ or opposite to it. This result has been obtained previously by Lipunov \& Shakura (1980) \footnote{Lipunov \& Shakura (1980) also argued that for $\cos^2\!\theta<1/3$, the minimum of the interaction energy between the central dipole and the field generated by the disk current is achieved at $\beta=0^\circ$, while for $\cos^2\!\theta>1/3$, the minimum energy corresponds to $\beta=90^\circ$. They therefore suggested that in the latter case ($\cos^2\!\theta>1/3$), the disk tends to evolve into the $\beta=90^\circ$ state. However, if eq.~(\ref{torque1}) is the only torque present in the disk, it is not clear how $\beta$ can change. Moreover, the ``magnetic Bardeen-Petterson'' effect always tends to align ${\hat l}$ and ${\hat\omega}$, independent of the sign of ${\Omega}_{\rm prec}$ (see \S 4.2). I became aware of the papers by Lipunov et al.~(see also Lipunov, Sem\"enov \& Shakura 1981) in mid-February 1999, at which point this paper was nearly finished. For completeness and pedagogical reason, I have decided to keep this subsection (\S 2.1) in its original form. I thank Dr.~Brad Hansen (CITA) for drawing my attention to the papers by Lipunov et al.}. \subsection{Magnetically Threaded Disk} We now consider the opposite limit in which the stellar magnetic field rapidly penetrates the disk (on a timescale shorter than the dynamical time of the disk) (see \S 2.3). Because of the shear between the disk and the plasma outside the disk, the threaded vertical field is winded to produce an azimuthal field which has different signs on the upper and lower surfaces of the disk. We thus adopt the following {\it ansatz} for the magnetic field in the disk\footnote{ In principle, an additional radial field $\Delta B_r$ could be generated (with opposite signs in the upper and lower disk surfaces) when the radial inflow of the disk drags the threaded $B_z$. This discontinuous $\Delta B_r$ will give rise to a precessional torque similar to that derived in \S 2.1. Since we expect $\Delta B_r$ to be proportional to the radial velocity, it is neglected in eq.~(\ref{br2}).}: \begin{eqnarray} B_r&=&{2\mu\over r^3}\sin\chi\cos(\phi-\phi_\mu),\label{br2}\\ B_\phi&=&{\mu\over r^3}\sin\chi\sin(\phi-\phi_\mu)\pm\zeta{\mu\over r^3}\cos\chi,\label{bphi2}\\ B_z&=&-{\mu\over r^3}\cos\chi.\label{bz2} \end{eqnarray} In (\ref{bphi2}), the second term represents the field produced by the twisting of $B_z$, and the upper (lower) sign corresponds to the value at the upper (lower) disk surface\footnote{ There could be an additional contribution to $B_\phi$ on the right-hand side of (\ref{bphi2}) due to the shearing of the stellar $B_r$ by the differential rotation. This contribution is not included for two reasons. First, the stellar $B_r$ may not be able to penetrate the thin disk; Second, even if a component of $B_r$ (e.g., the static component; see \S 2.3) penetrates the disk, the induced $\Delta B_\phi$ is expected to be smaller than that due to the threaded $B_z$. To see this, we write, schematically, $${\partial \Delta B_\phi\over \partial t}=r\left(B_z{\partial\Omega\over\partial z}+B_r {\partial\Omega\over\partial r}\right)-{\Delta B_\phi\over\tau_{\rm diss}},$$ where $B_z$ and $B_r$ are the threaded fields, $\tau_{\rm diss}$ is the effective dissipation time. In steady state, we find $$\Delta B_\phi=\mp \zeta B_z-{3\over 2} \tau_{\rm diss}\Omega B_r,$$ where $\zeta=(r/H_B)\tau_{\rm diss} (\Omega-\omega)$ (for closed field configurations) or $\zeta=(r/H_B)\tau_{\rm diss}\Omega$ (for open field configurations) ($H_B$ is the vertical scale in which $\Omega$ varies). Note that the shearing of $B_r$ does not produce any surface current. For $H_B\ll r$, we can drop the term proportional $B_r$ in $\Delta B_\phi$, and obtain $\Delta B_\phi=\mp \zeta B_z$. }. The quantity $\zeta$ specifies the azimuthal pitch of the field line. In general we expect $|\zeta|\lo 1$ (e.g., Sturrock \& Barnes 1972; Lovelace et al.~1995 and references therein), but its actual value or form depends on details of the dissipative processes involved in the disk-magnetic field interactions. If the stellar magnetic field threads the disk in a closed configuration (e.g., Ghosh \& Lamb 1979a,b; Wang 1987,1995), we expect $\zeta\propto (\Omega-\omega)$ so that $\zeta>0$ for $\Omega>\omega$ and $\zeta<0$ for $\Omega<\omega$. But it has been argued that the differential shearing and the plasma flowing from the disk into the overlying magnetosphere will blow the field lines open and maintain them in a open configuration (e.g., Arons 1987; Newman et al.~1992; Lynden-Bell \& Boily 1994; Lovelace et al.~1995 and references therein), in which case we expect $\zeta$ to be positive and of order unity. For our purpose in this paper, $\zeta$ is simply a dimensionless number or function\footnote{Even in models (e.g., Shu et al.~1994) where the disk is largely diamagnetic with no intrinsic $B_z$, the stellar field must penetrate the boundary layer. In this case, one would imagine that $\zeta$ is nonzero only near the magnetosphere boundary.}. Also note that because of the screening currents in the disk (such as those discussed in \S 2.1) or in the magnetosphere boundary, the threaded field may be smaller than the vacuum field produced by the dipole. But since our result (see below) depends on $\zeta\mu^2$, we can easily absorb the screening effect into the definition of $\zeta$. The surface current on the disk corresponding to the field (\ref{br2})-(\ref{bz2}) is \begin{equation} {\bf K}=-{\zeta c\mu\over 2\pi r^3}\cos\chi\,{\hat r}. \end{equation} The interaction of ${\bf K}$ with $B_z$ gives rise to an azimuthal force, which tends to slow down (or speed up if $\omega>\Omega$ and the field lines are in a closed configuration) the fluid motion, transferring angular momentum between the disk and the overlying magnetosphere or the star. This is the familiar magnetic breaking torque which has been included in previous studies of aligned magnetized disk (e.g., Ghosh \& Lamb 1979a,b; Lovelace et al.~1995; Wang 1995; Lai 1998). Here we are interested in the vertical magnetic force $F_z$ (per unit area), resulting from the interaction between ${\bf K}$ and the azimuthal field produced by the stellar dipole (the first term in [\ref{bphi2}]): \begin{equation} F_z=-{\zeta\mu^2\over 4\pi r^6}\sin 2\chi\sin(\phi-\phi_\mu). \label{force2}\end{equation} This is simply the difference in $B^2/(8\pi)$ below and above the disk. The magnetic torque associated with $F_z$ is given by \begin{equation} {\bf N}={\zeta\mu^2\over 4\pi r^5}\sin 2\chi \sin (\phi-\phi_\mu)\,\hat\phi. \end{equation} Averaging over $\phi$, we have \begin{equation} \langle{\bf N}\rangle_\phi=-{\zeta\mu^2\over 8\pi r^5}\sin 2\chi (\cos\phi_\mu{\hat x}+\sin\phi_\mu{\hat y}) =-{\zeta\mu^2\over 4\pi r^5}({\hat l}\cdot{\hat\mu})\left[{\hat\mu}-({\hat\mu}\cdot {\hat l}){\hat l}\right]. \label{torquephi1}\end{equation} For a nonrotating star, this torque tends to pull the disk normal vector ${\hat l}$ toward being perpendicualr to the magnetic axis ${\hat\mu}$ (assuming $\zeta>0$), thus making the disk plane parallel to the stellar field lines. For a rotating star, averaging over the spin period and using the indentities (\ref{id1})-(\ref{id3}), we find \begin{eqnarray} \langle\langle{\bf N}\rangle\rangle &=& -{\zeta\mu^2\over 16\pi r^5}\sin 2\beta \left(3\cos^2\theta-1\right)\,{\hat y} \nonumber\\ &=&-{\zeta\mu^2\over 8\pi r^5}\cos\beta \left(3\cos^2\theta-1\right) \left[{\hat\omega}-({\hat\omega}\cdot{\hat l}){\hat l}\right]. \label{torque2}\end{eqnarray} In the absence of other forces, this magnetic torque will change the tilt angle $\beta$ of the ring (at radius $r$) according to \begin{equation} {d\beta\over dt}={\zeta\mu^2\over 16\pi r^7\Omega\Sigma}\sin 2\beta \left(3\cos^2\theta-1\right). \label{dbeta1}\end{equation} Thus, depending on the sign of $\zeta\cos\beta (3\cos^2\theta-1)$, the angle $\beta$ can increase or decrease: For $\zeta (3\cos^2\theta-1)>0$, the torque drives the disk toward the perpendicular configuration ($\beta=90^\circ$), while for $\zeta (3\cos^2\theta-1)<0$, it drives the disk toward alignment ($\beta=0^\circ$) or anti-alignment ($\beta=180^\circ$). In particular, when $\zeta (3\cos^2\theta-1)>0$, the aligned configuration is unstable against the growth of disk tilt angle --- This is the {\it magnetic warping instability} (see \S 2.3 and \S 2.4). It is instructive to understand the difference between the situation studied here and that of a diamagnetic disk considered in \S2.1. In both cases, a vertical magnetic force is exerted on the disk. Although eqs.~(\ref{force1}) and (\ref{force2}) appear similar, they have very different spin-averaged behavior. For a diamagnetic disk, we have \begin{equation} \langle F_z\rangle_{\omega}={\mu^2\over\pi^2 r^6D}\sin 2\beta\left(3\cos^2\theta-1\right)\,\sin\phi. \end{equation} Thus the force on the $y>0$ side of the disk has a different sign from that on the $y<0$ side, giving rise to a torque which is along the $x$-axis (perpendicular to both ${\hat l}$ and ${\hat\omega}$). For a magnetically threaded disk, we have \begin{equation} \langle F_z\rangle_{\omega}={\zeta\mu^2\over 8\pi r^6}\sin 2\beta\left(3\cos^2\theta-1\right)\,\cos\phi. \end{equation} The force has different signs for $x>0$ and $x<0$, and the torque is along the $y$-axis (in the same plane as ${\hat l}$ and ${\hat\omega}$). \subsection{A Hybrid ``Realistic'' Model} The results of \S 2.1 and \S 2.2 represent two opposite, extreme situations, and are not likely to be realistic. The magnetic field configuration in a purely diamagnetic disk (see \S 2.1) is prone to Kelvin-Helmholtz instability and reconnection. It has been argued that the non-linear development of the these processes could lead to partial threading of the magnetic field through the disk (e.g., Ghosh \& Lamb 1979a,b; Wang 1987,~1995), although the extent of the field-threading region is uncertain and the closed configuration advocated by Ghosh and Lamb is probably unrealistic (e.g., Arons 1987; Shu et al.~1994; Lovelace et al.~1995). In any event, near the magnetosphere boundary, the vertical stellar field must certainly thread the disk plasma. On the other hand, the magnetically thread disk model considered in \S 2.2 requires that the stellar field penetrates the disk almost instantaneously (compared to the spin period and the orbital period). This is unrealistic. The timescale for field penetration is uncertain, but it cannot be shorter than the dynamical time of the disk; it may be as long as the disk thermal time (E.~T. Vishniac 1999, private communication; see Park \& Vishniac 1996 and references therein). The vertical vacuum field produced by the stellar dipole on the disk can be written as the sum of a static component and a time-varying component: \begin{equation} B_z^{(0)}=-{\mu\over r^3}\cos\chi=-{\mu\over r^3}\cos\beta\cos\theta +{\mu\over r^3}\sin\beta\sin\theta\sin\omega t. \label{bz0}\end{equation} While it is possible that the static field can penetrate the disk by allowing sufficient time for Kelvin-Helmholtz instability/reconnection to grow, it is almost certain that the variable field will be shielded out of the disk by the screening current. We therefore consider the following hybrid model, which we consider to be more realistic than the situations studied in \S 2.1 and \S 2.2: The static component of the vertical stellar field threads the disk, while the time-varying component is screened out by the disk. The winding of the threaded field will produce an azimuthal field and a radial surface current, while the variable vertical field will induce a shielding azimuthal surface current and discontinuous radial field --- this radial field can be obtained by appropriately modifying Aly's solution as given in \S 2. The magnetic field on the disk is then given by \footnote{In eq.~(\ref{br3}) we have neglected a possible component of $B_r$ generated by radial infall of the threaded $B_z$; See footnote 2. Also, in (\ref{bphi3}) we have ignored a possible toroidal field generated by the shearing of the stellar $B_r$; see footnote 3.} \begin{eqnarray} B_r&=&{2\mu\over r^3}\sin\chi\cos(\phi-\phi_\mu)\pm{4\mu\over\pi r^3D}\sin\beta\sin\theta\sin\omega t,\label{br3}\\ B_\phi&=&{\mu\over r^3}\sin\chi\sin(\phi-\phi_\mu)\pm\zeta{\mu\over r^3}\cos\beta\cos\theta,\label{bphi3}\\ B_z&=&-{\mu\over r^3}\cos\beta\cos\theta.\label{bz3} \end{eqnarray} In (\ref{br3}), $B_r$ is the sum of the vacuum dipole field and the field produced by the azimuthal screening current (which is induced to cancel the variable part of $B_z^{(0)}$); in (\ref{bphi3}), $B_\phi$ is the sum of the vacuum dipole field and the field induced by the winding of the constant $B_z$ (see \S 2.2 for the property of $\zeta$). The surface current on the disk is \begin{equation} {\bf K}={2c\mu\over\pi^2r^3D}\sin\beta\sin\theta\sin\omega t\,\hat\phi- {\zeta c\mu\over 2\pi r^3}\cos\beta\cos\theta\,\,\hat r. \end{equation} The vertical magnetic force (per unit area) on the disk is then given by \begin{equation} F_z=-{4\mu^2\over\pi^2 r^6D}\sin\beta\sin\theta \sin\omega t\sin\chi\cos(\phi-\phi_\mu) -{\zeta\mu^2\over 2\pi r^6}\cos\beta\cos\theta\sin\chi\sin(\phi-\phi_\mu). \label{force3}\end{equation} The corresponding torque (per unit area) acting on the disk, averaged over the ring, is given by \begin{equation} \langle{\bf N}\rangle_\phi=-{2\mu^2\over\pi^2 r^5D}\sin\beta\sin\theta \sin\omega t\,({\hat\mu}\times{\hat l}) -{\zeta\mu^2\over 4\pi r^5}\cos\beta\cos\theta\,\left[{\hat\mu} -({\hat\mu}\cdot{\hat l}){\hat l}\right]. \label{torquephi}\end{equation} For nonrotating star, eq.~(\ref{torquephi}) reduces to eq.~(\ref{torquephi1}) (note that $\cos\chi=\cos\beta\cos\theta$ when $\omega=0$): The torque tends to make the disk plane align with the stellar field lines (i.e., ${\hat l}$ perpendicular to ${\hat\mu}$). Averaged over the stellar rotation, the torque can be written as \begin{equation} \langle\langle{\bf N}\rangle\rangle= \langle\langle{\bf N}\rangle\rangle_{\rm prec}+ \langle\langle{\bf N}\rangle\rangle_{\rm warp}, \label{torque3a}\end{equation} where the precessional torque and the warping torque are given by \begin{eqnarray} \langle\langle{\bf N}\rangle\rangle_{\rm prec} &=&-{\mu^2\over\pi^2 r^5D}\cos\beta\sin^2\!\theta\,({\hat\omega}\times{\hat l}),\label{torque3b}\\ \langle\langle{\bf N}\rangle\rangle_{\rm warp} &=&-{\zeta\mu^2\over 8\pi r^5}\sin 2\beta\cos^2\!\theta\,{\hat y} =-{\zeta\mu^2\over 4\pi r^5}\cos\beta\cos^2\!\theta\,\left[ {\hat\omega}-({\hat\omega}\cdot{\hat l}){\hat l}\right]. \label{torque3c}\end{eqnarray} Clearly, the diamagnetic feature of the disk induces a precessional torque (eq.~[\ref{torque3b}]), with the precession angular frequency given by \begin{equation} {\bf\Omega}_{\rm prec}=-{\mu^2\over\pi^2 r^7\Omega\Sigma D}\cos\beta\sin^2\!\theta\,{\hat\omega}. \label{prec3}\end{equation} On the other hand, the disk-threading field gives rise to a torque (eq.~[\ref{torque3c}]) which changes $\beta$ at a rate given by \begin{equation} {d\beta\over dt}={\zeta\mu^2\over 8\pi r^7\Omega\Sigma} \sin 2\beta\cos^2\!\theta. \label{warp3}\end{equation} Note that the disk precession is along the direction $(-\cos\beta\,{\hat\omega})$, i.e., the precession is always in the opposite sense as the orbital motion of disk. Equation (\ref{warp3}) reveals the {\it magnetic warping instability}: the spin-orbit inclination angle $\beta$ increases when $\zeta\cos\beta>0$ and decreases when $\zeta\cos\beta>0$. Thus, for $\zeta>0$ (the most likely case; see \S 2.2), the warping torque always tends to drive the disk toward the configuration where ${\hat l}$ is perpendicular to ${\hat\omega}$. \subsection{General Consideration} The preceding subsections assume that the vertical field that threads the disk originates from the star. But it is important to note that this is not a requirement for the existence of the warping torque. Indeed, models of hydromagnetic driven outflows from disks are predicated on the existence of large-scale poloidal magnetic field that threads the disk --- this field could have been advected inward by the accreting flow or generated in the disk (see, e.g., Blandford 1989 for a review). Suppose there is vertical $B_z$ (assumed to be time-independent) which threads the disk. The warping torque results from the interaction of the stellar $B_\phi$ and the radial surface current on the disk induced by the twisting of the threaded $B_z$. Thus in general, we can write the azimuthal field on the disk as \begin{equation} B_\phi={\mu\over r^3}\sin\chi\sin(\phi-\phi_\mu)\mp\zeta B_z, \end{equation} with $\zeta$ a positive dimensionless number of order or less than unity. The vertical force on the disk is given by \begin{equation} F_z={\zeta\mu B_z\over 2\pi r^3}\sin\chi\sin(\phi-\phi_\mu). \label{force4}\end{equation} The averaged (warping) torque is then \begin{equation} \langle\langle{\bf N}\rangle\rangle={\zeta\mu B_z\over 4\pi r^2}\sin\beta \cos\theta\,{\hat y}. \label{torque4}\end{equation} The model considered in \S 2.3 corresponds to $B_z=-\mu\cos\beta\cos\theta/r^3$. The precessional torque results from the interaction of the ring of diamagnetic screening current and the radial magnetic field produced by the star. Assume that the surface current has the form \begin{equation} {\bf K}=\left(K_{\phi 1}+K_{\phi 2}\sin\omega t\right)\,\hat\phi. \end{equation} (In general, both the static current $K_{\phi 1}$ and the time-dependent current $K_{\phi 2}\sin\omega t$ are possible as the disk responds to the external field that tries to enters the disk.) The vertical magnetic force is \begin{equation} F_z=-{2\mu\over cr^3}\left(K_{\phi 1}+K_{\phi 2}\sin\omega t\right)\sin\chi \cos(\phi-\phi_\mu), \label{force5}\end{equation} and the resulting averaged (precessional) torque is \begin{equation} \langle\langle{\bf N}\rangle\rangle=-{\mu\over cr^2}\left(K_{\phi 1}\sin\beta \cos\theta+{K_{\phi 2}\over 2}\cos\beta\sin\theta\right)\,{\hat x}. \label{torque7}\end{equation} The model considered in \S 2.3 corresponds to $K_{\phi 1}=0$ and $K_{\phi 2}=(2c\mu/\pi^2 r^3D)\sin\beta\sin\theta$, while in the model considered in \S 2.1, we have $K_{\phi 1}=-(2c\mu/\pi^2r^3D) \cos\beta\cos\theta$ and $K_{\phi 2}=(2c\mu/\pi^2 r^3D)\sin\beta\sin\theta$. Finally, we note that in our calculations of the warping torques (\S 2.2-\S2.4) we have assumed that the azimuthal pitch $\zeta$ is stationary in time. An alternative scenario has been proposed (Aly \& Kuijpers 1990; van Ballegooijen 1994), where the threaded field lines are being constantly wound up, with occasional field reconnection which releases the stored-up magnetic energy. Even for such a time-dependent magnetic field structure (apart form the time-dependence of the rotating dipole field), it is likely that magnetic torques (similar to those calculated in this paper) will exist, although this possible complication will be ignored in this paper. \subsection{Disk Consisting of Diamagnetic Blobs} The previous subsections treat the disk as a continuum. It has been suggested that under certain conditions, the accretion disk may be lumpy, consisting of diamagnetic blobs (Vietri \& Stella 1998; see King 1993). Such inhomogeneous accretion flow may arise from the nonlinear development of various plasma instabilities (e.g., Kelvin-Helmholtz instability; Arons \& Lea 1980) associated with the disk and the magnetosphere. Vietri \& Stella (1998) studied the motion of diamagnetic blobs orbiting a central star with an inclined magnetic dipole and showed that magnetic drag force (Drell, Foley \& Ruderman 1965) on the blob can induce vertical resonances near the corotation radius of the disk. We have calculated the magnetic torque on the accretion disk consisting of individual diamagnetic blobs. It can be shown that independent of the vertical resonances, there exists a magnetic torque which tends to induce tilt on the orbit of the blob. Since there are considerable uncertainties associated with such an extreme form of accretion disk, we relegate the calculation to Appendix C. \section{MAGNETICALLY DRIVEN RESONANCES} The discussion in \S 2 has neglected possible resonances in the interaction between the disk the rotating central dipole. Here we consider these resonances. \subsection{Vertical and Epicyclic Resonances} The bending mode of the disk is characterized by a small vertical displacement $Z(r,\phi,t)$, with the equation of motion \begin{equation} {d^2Z\over dt^2}= \left({\partial\over\partial t}+\Omega{\partial\over\partial\phi}\right)^2Z =-\Omega_z^2 Z+{1\over\Sigma}F_z, \label{eom1}\end{equation} where $\Omega$ is the orbital angular frequency, the term $-\Omega_z^2z$ represents vertical the gravitational restoring force (for a Keplerian disk, $\Omega_z=\Omega$), and $F_z$ is the magnetic force (per unit area) as calculated in \S 2. Pressure, viscous force and self-gravity have been neglected. Consider first the magnetic force given by eq.~(\ref{force4}), and write it as \begin{equation} F_z={\zeta\mu B_z\over 2\pi r^3}\Bigl[\sin\theta\cos\omega t\sin\phi -\left(\sin\beta\cos\theta+\cos\beta\sin\theta\sin\omega t\right) \cos\phi\Bigr]. \label{force6}\end{equation} There exists the following possible vertical resonances: \begin{eqnarray} &&\Omega_z=\Omega\quad\qquad\qquad~~~~~ ({\rm for }~~\sin\beta\cos\theta\neq 0)\label{res1}\\ &&\omega-\Omega=\pm\Omega_z\qquad\qquad({\rm for }~~\sin\theta\neq 0~~ {\rm and}~~\beta\neq \pi)\label{res2}\\ &&\omega+\Omega=\Omega_z\qquad\qquad~~ ({\rm for }~~\sin\theta\neq 0~~{\rm and} ~~\beta\neq 0).\label{res3} \end{eqnarray} These resonances can be easily understood: (1) When $\sin\beta\cos\theta\neq 0$, there exists a static distribution ($\propto \cos\phi$) of vertical force field, and a fluid element feels the force $\propto\cos\Omega t$; this explains the $\Omega=\Omega_z$ resonance. (2) When $\sin\theta\neq 0$, there exists a time-dependent vertical force field ($\propto\sin\omega t$ or $\cos\omega t$). A fluid element, traveling with angular velocity $\Omega$, would experience vertical forces with frequency $(\omega\pm\Omega)$ [for $\beta=0$, only $(\omega-\Omega)$ is possible, while for $\beta=180^\circ$, only $(\omega+\Omega)$ is possible]; this explains the $\omega\pm\Omega=\pm\Omega_z$ resonances. Now consider the magnetic force given by eq.~(\ref{force5}). The force associated with $K_{\phi 1}$ has the form $\sin\phi$ and $\cos(\phi\pm\omega t)$, which would give rise to the $\Omega_z=\Omega$ resonance and the $\omega\pm\Omega=\pm\Omega_z$ resonances as in eqs.~(\ref{res1})-(\ref{res3}). The force associated with $K_{\phi 2}$ has a time-dependence of the form \begin{eqnarray} \sin\omega t\sin\chi\cos(\phi-\phi_\mu)&=& {1\over 2}\cos\beta\sin\theta\sin\phi+ \sin\beta\cos\theta\sin\omega t\sin\phi\nonumber\\ &&+{1\over 2}\sin\theta\Bigl(\sin 2\omega t\cos\phi-\cos\beta\cos 2\omega t \sin\phi\Bigr). \end{eqnarray} This gives rise to the following resonances: \begin{eqnarray} &&\Omega_z=\Omega\quad\qquad\qquad~~~~~~~\, ({\rm for }~~\cos\beta\sin\theta\neq 0)\label{reso1}\\ &&\omega\pm \Omega=\pm\Omega_z\qquad\qquad~~\,({\rm for }~~\sin\beta\cos\theta \neq 0)\label{reso2}\\ &&2\omega-\Omega=\pm\Omega_z\qquad\qquad~ ({\rm for }~~\sin\theta\neq 0~~{\rm and}~~\beta\neq \pi)\label{reso3}\\ &&2\omega+\Omega=\Omega_z\qquad\qquad~~~ ({\rm for }~~\sin\theta\neq 0~~{\rm and}~~\beta\neq 0).\label{reso4} \end{eqnarray} The new resonances, $2\omega\pm\Omega=\pm\Omega_z$, come about because the magnetic field varies as $\cos\omega t$ or $\sin\omega t$, and the screening current also varies as $\cos\omega t$ or $\sin\omega t$. In addition to the vertical resonances discussed above, epicyclic resonance can also arise from the magnetic field -- disk interaction. Consider the field given by eqs.~(\ref{br3})-(\ref{bz3}). The interaction between $K_\phi$ with $B_z$ gives rise to a radial force \begin{equation} F_r=-{\mu^2\over 2\pi^2r^6D}\sin 2\beta\sin 2\theta\sin\omega t. \end{equation} Clearly, epicyclic resonance occurs when \begin{equation} \kappa=\omega\qquad\qquad({\rm for}~~\sin 2\beta\sin 2\theta\neq 0), \label{epy}\end{equation} where $\kappa$ is the epicyclic frequency. For a Keplerian disk, $\kappa=\Omega$, (\ref{epy}) is a corotation resonance. We note that this paper deals only with dipole field from the star. When higher-order multipole fields are considered, it is conceivable that additional resonances can arise. \subsection{Magnetic Torques at Resonances} What are the consequences of the magnetically driven resonances? We have not investigated this issue in detail. The resonances may act as an extra source (in addition to the non-resonant warping torques discussed in \S 2 and \S 4) for generating bending waves and spiral waves in the disk. Near the resonances, fluid elements undergo large out-of-plane and radial excursions, which may lead to thickening of the disk\footnote{This may be analogous to the Lorentz resonances (which occur when charge particles move around a rotating magnetic field) in the jovian ring (Burns et al.~1985; Schaffer \& Burns 1992). However, because of the fluid nature of the disk, the resonances may not lead to sharp edges in the disk.}. We can get some insight into the resonances by calculating the magnetic torques at the resonant radii. Consider first the torque associated with the force in eq.~(\ref{force6}). For a Keplerian disk, the condition $\Omega_z=\Omega$ is always satisfied. When $\sin\beta\cos\theta\neq 0$, there always exists a net (averaged) torque acting on the fluid element, pulling its orbital angular momentum axis toward (or away from) the spin axis (eq.~[\ref{torque4}]). The condition $\Omega+\Omega_z=\omega$ (or $2\Omega=\omega$ for a Keplerian disk) specifies a special radius in the disk. To calculate the net torque on the fluid element at this resonant radius, we cannot average $\phi$ and the spin period independently (as done in \S 2). Instead we write the torque ${\bf N}={\bf r}\times{\bf F}$ as: \begin{eqnarray} {\bf N}&=&{\zeta\mu B_z\over 4\pi r^2}\biggl\{ \Bigl[\sin\theta\cos\omega t (1-\cos 2\phi)-(\sin\beta\cos\theta+\cos\beta \sin\theta\sin\omega t)\sin 2\phi\Bigr]\,{\hat x}\nonumber\\ &&+\Bigl[-\sin\theta\cos\omega t\sin 2\phi+(\sin\beta\cos\theta +\cos\beta\sin\theta\sin\omega t)(1+\cos 2\phi)\Bigr]\,{\hat y}\biggr\}. \end{eqnarray} Following a fluid element we have $\phi=\Omega t+\phi_0$. Averaging over time, we find that, at the $2\Omega=\omega$ resonance, the torque is given by \begin{equation} \langle\langle{\bf N}\rangle\rangle={\zeta\mu B_z\over 4\pi r^2}\sin\beta \cos\theta\,{\hat y} -{\zeta\mu B_z\over 8\pi r^2}(1+\cos\beta)\sin\theta \,({\hat x}\cos 2\phi_0+{\hat y}\sin 2\phi_0). \label{torque5}\end{equation} Thus for each fluid element on the resonant radius, the averaged torque is modified from the ``nonresonant'' value (the first term in eq.~[\ref{torque5}]). Since different fluid elements in the same ring have different values of $\phi_0$, they will experience different torques; if they are allowed to move independent of each other, the ring will eventually disperse. However, if there is strong coupling between the different elements on the ring so that the ring evolve dynamically as an identity, then one should average over $\phi_0$, and eq.~(\ref{torque5}) reduces to (\ref{torque4}). We can similarly consider the torque associated with the magnetic force given in eq.~(\ref{force5}). At both the $\Omega=\omega$ and $2\Omega=\omega$ resonances, in addition to eq.~(\ref{torque7}), there are ``resonant'' torques which depend on $\phi_0$ (analogous to eq.~[\ref{torque5}]). We note that the ``nonresonant'' torque is nonzero only when $\sin\beta\neq 0$ (misaligned spin-orbit), while at the $2\Omega=\omega$ or $\Omega=\omega$ resonances, the torque can be nonzero even when $\beta=0$. Also note that in general, the ``resonant'' torque is of the same order of magnitude as the ``nonresonant'' torque. We shall not consider these resonances in the rest of the paper. \section{DYNAMICS OF WARPING AND PRECESSING DISK} We now study the dynamics of the disk under the influence of the magnetic torques calculated in \S 2. For concreteness, we shall use the torque expressions of \S 2.3; using other expressions of \S 2 would give similar results (although the dependence on angles would be different). We are particularly interested in whether the warping instability can operate in the presence of disk viscosity. \subsection{Criterion for the Warping Instability} Our starting point is the evolution equation for disk tilt ${\hat l} (r,t)$ (the unit vector perpendicular to the disk annulus at radius $r$) given by Pringle (1992) (see also Papaloizou \& Pringle 1983) \footnote{Recent study (Ogilvie 1999) indicates that in the nonlinear regime, the warping equation needs to be modified. In effect, $\nu_1$ and $\nu_2$ are not the usual vertically averaged viscosities and may depend on the amplitude of the warp.}: \begin{eqnarray} &&{\partial{\hat l}\over\partial t}+\left[V_r-{\nu_1\Omega'\over\Omega} -{1\over 2}\nu_2{(\Sigma r^3\Omega)'\over\Sigma r^3\Omega}\right] {\partial{\hat l}\over\partial r}\nonumber\\ &&\quad\qquad ={\partial\over\partial r}\left({1\over 2}\nu_2{\partial{\hat l} \over\partial r}\right)+{1\over 2}\nu_2\left|{\partial{\hat l}\over\partial r} \right|^2+{{\bf N}\over\Sigma r^2\Omega}. \label{warp}\end{eqnarray} Here $V_r$ is the radial velocity of the flow, $\Omega'\equiv d\Omega/dr$, $\nu_1$ is the usual disk viscosity (measuring the $r$-$\phi$ stress), and $\nu_2$ is the viscosity (measuring the $r$-$z$ stress) associated with reducing disk tilt. We assume that the timescale for ${\hat l}$ to change is much longer than the spin period, so that ${\bf N}$ is the averaged torque (per unit area) as calculated in \S 2.3 (The notation $\langle\langle\cdots\rangle\rangle$ has been suppressed). Using the following relations for a Keplerian disk: \begin{equation} V_r=-{3\nu_1\over 2r}{\cal J}^{-1},~~~~ \Sigma={\dot M\over 3\pi\nu_1}{\cal J}, \label{vr}\end{equation} where ${\cal J}$ is a function of $r$ which approaches unity in the region far from the inner edge of the disk (see Appendix A), and assuming that $\nu_2/\nu_1$ is independent of $r$, we reduce eq.~(\ref{warp}) to \begin{eqnarray} &&{\partial{\hat l}\over\partial t}-\left[{3\nu_2\over 4r}\left(1+ {2r{\cal J}'\over 3{\cal J}}\right)+{3\nu_1\over 2r}\left({\cal J}^{-1}-1\right)\right] {\partial{\hat l}\over\partial r}\nonumber\\ &&\quad\qquad ={1\over 2}\nu_2{\partial^2{\hat l}\over\partial r^2} +{1\over 2}\nu_2\left|{\partial{\hat l}\over\partial r} \right|^2+{{\bf N}\over\Sigma r^2\Omega}. \label{warp1}\end{eqnarray} We rewrite the magnetic torques in eqs.~(\ref{torque3b}), (\ref{torque3c}) as \begin{eqnarray} {{\bf N}_{\rm prec}\over\Sigma r^2\Omega}&=&-\Omega_p\cos\beta\,\,{\hat\omega}\times{\hat l}, \,~~~~~~~~~\Omega_p={\mu^2\over\pi^2r^7\Omega\Sigma D}\sin^2\theta, \label{omegap} \\ {{\bf N}_{\rm warp}\over\Sigma r^2\Omega}&=&-\Gamma_w\cos\beta\sin\beta\,{\hat y}, ~~~~~~~~\Gamma_w={\zeta\mu^2\over 4\pi r^7\Omega\Sigma}\cos^2\theta.\label{gammaw} \end{eqnarray} At this stage ${\hat l}$ is no longer useful as a coordinate axis, so we consider a different cartesian coordinate system where ${\hat\omega}$ is the $z'$-axis (see Fig.~1). In this coordinate system, we write \begin{equation} {\hat l}=(\sin\beta\cos\gamma,\sin\beta\sin\gamma,\cos\beta), \end{equation} which defines the local twist angle $\gamma$. For small tilt angle ($\beta\ll 1$), eq.~(\ref{warp1}) simplifies to \begin{equation} {\partial W\over\partial t}-{3\nu_2\over 4r}{\partial W\over\partial r} ={1\over 2}\nu_2{\partial^2W\over\partial r^2} -i\Omega_pW+\Gamma_wW, \label{warpw}\end{equation} where $W\equiv \beta\, {\rm e}^{i\gamma}$, and we have set ${\cal J}=1$ (far from the inner edge of the disk) -- using a more rigorous ${\cal J}$ would only affect the numerical coefficient in front of the $\partial W/\partial r$ term. To derive the instability criterion, we consider the WKB solution of the form (valid for $kr\gg 1$) \begin{equation} W\propto \exp(i\sigma t+ikr). \end{equation} Substituting into (\ref{warpw}), we obtain the dispersion relation \begin{equation} \sigma=\left({3\nu_2\over 4r}k-\Omega_p\right)+i\left({1\over 2} \nu_2 k^2-\Gamma_w\right). \end{equation} For the instability to grow, we require ${\rm Im}(\sigma)<0$, or \begin{equation} \Gamma_w>{1\over 2}\nu_2 k^2 \Longleftrightarrow {\rm Instability}. \end{equation} Since the radial wavelength is restricted to $\lambda\le r$, or $k\ge 2\pi/r$, the instability criterion becomes \begin{equation} \Gamma_w >{2\pi^2}{\nu_2\over r^2}\Longleftrightarrow {\rm Instability}. \label{criterion}\end{equation} This criterion has a simple physical interpretation: The magnetic torque drives the growth of disk tilt on a timescale $\Gamma_w^{-1}$, while viscosity tries to reduce the tilt on a timescale $r^2/\nu_2$ (which is of the same order as the radial drift time of the flow, $r/|V_r|$, if $\nu_2$ is of the same order as $\nu_1$). Instability requires $\Gamma_w^{-1}\lo r^2/\nu_2$. Using (\ref{vr}) and (\ref{gammaw}), and assuming that $\nu_2/\nu_1$ is independent of $r$, we can further reduce (\ref{criterion}) to \begin{equation} r<r_w=\left({3\,\zeta\cos^2\theta\over 8\pi^2{\cal J}}{\nu_1\over\nu_2} \right)^{2/7}\left({\mu^4\over GM\dot M^2}\right)^{1/7}. \end{equation} This indicates that inside the critical {\it warping radius} $r_w$, the magnetic torque can overcome the viscous force and make the disk tilt grow. Moreover, with the expected\footnote{For Keplerian and inviscid (or very nearly so) disks, resonance between the epicyclic frequency and orbital frequency leads to $\nu_2/\nu_1\simeq 1/(2\alpha^2)\gg 1$. However, the resonance is delicate and might be destroyed by the effects of general relativity, self-gravity, magnetic fields and turbulence. See Ogilvie (1999) for a discussion.} $\zeta\sim 1$, $\nu_1/\nu_2\sim 1$ and ${\cal J}\le 1$, the warping radius is of the same order of magnitude as the magnetosphere radius $r_m$ (see eq.~[\ref{alfven}] and Appendix B for a discussion of the magnetosphere radius for arbitrary geometry). Thus, as the disk approaches the magnetosphere, it will tend to be tilted with respect to the stellar spin even if at large radii the disk normal is aligned with the spin axis. Similar analysis can be applied to the case where the disk is nearly antiparallel to the spin ($\pi-\beta\ll 1$). If we define $W\equiv (\pi-\beta){\rm e}^{i\gamma}$, then eq.~(\ref{warpw}) remains valid except the sign in front of $i\Omega_pW$ changes. We obtain the same instability criterion as above. Thus, an antiparallel disk tends to be driven toward a perpendicular configuration near the magnetosphere radius. \subsection{Magnetic Bardeen-Petterson Effect} A tilted disk will be driven into precession by the torque ${\bf N}_{\rm prec}$ with ${\bf\Omega}_{\rm prec}=-\Omega_p\cos\beta\,{\hat\omega}$ (see eq.~[\ref{omegap}]). What is the effect of this precession on the disk tilt? An analogy can be made with the behavior of a disk undergoing Lense-Thirring precession around a rotating black hole (or any compact object) (Bardeen \& Petterson 1975; Kumar \& Pringle 1985; Pringle 1992; Scheuer \& Feiler 1996). The gravitomagnetic force from the rotating object (with spin angular momentum ${\bf J}=J{\hat\omega}$) drives the precession of the misaligned disk at angular frequency ${\bf\Omega}_{\rm LT}={2G{\bf J}/(c^2r^3)}$. Bardeen \& Petterson (1975) pointed out that the action of viscosity on the differentially precessing disk tends to align the rotation of the inner disk with the spin axis of the central object. The Bardeen-Petterson radius $r_{\rm BP}$, that is, loosely speaking, the radius inside which the disk aligns with the spin, is obtain by equating the precessional time $\Omega_{\rm LT}^{-1}$ and the viscous time $r^2/\nu_2$ (Scheuer \& Feiler 1996), i.e., \begin{equation} r_{\rm BP}={2GJ\over c^2\nu_2}. \end{equation} However, the transition from the warped outer disk to the aligned inner disk is rather broad, and for this reason the timescale to achieve the Bardeen-Petterson alignment is much larger than the precession timescale at $r_{\rm BP}$ (Kumar \& Pringle 1985; Pringle 1992). Numerical simulations (Pringle 1992) indicate that a steady state is achieved on a time scale of order $(10-100)\,\Omega_{\rm LT}^{-1}$ (evaluated at $r_{\rm BP}$). Now consider the effect of magnetically driven precession\footnote{In some systems (such as accreting neutron stars with weak magnetic fields), the Lense-Thirring precession dominates (See \S 6). But here we neglect $\Omega_{\rm LT}$.}. Setting $\Omega_p$ equal to $\nu_2/r^2$, we obtain the magnetic Bardeen-Petterson radius: \begin{equation} r_{\rm MBP}=\left({3\sin^2\theta\over\pi{\cal J} D}{\nu_1\over\nu_2}\right)^{2/7} \left({\mu^4\over GM\dot M^2}\right)^{1/7}. \end{equation} Inside $r_{\rm MBP}$, the combined effect of viscosity and precession tends to align the disk normal with the spin axis. We see that typically $r_{\rm MBP}$ is of the same order as $r_w$ (the warping radius) and $r_m$ (the magnetosphere radius). Thus the precessional torque has an opposite effect on the disk tilt as the warping torque discussed in \S 5.1. However, because of the broad warp-alignment transition expected for the magnetic Bardeen-Petterson effect and the long timescale involved, we expect that the precession-induced alignment will be overwhelmed by the warping instability. \section{EFFECT ON THE SPIN EVOLUTION} How does the magnetically warped and precessing disk affect the spin evolution of the central star? The angular momentum of the accreting gas is deposited at the magnetospheric boundary and transferred to the the central object. In addition, there are back-reactions on the star associated with the magnetic torques on the disk. Thus the spin angular velocity {\boldmath $\omega$} evolves according to \begin{equation} {d\over dt}\left(I{\mbox{\boldmath $\omega$}}\right)=\dot M (GMr_m)^{1/2}{\hat l} -{\mbox{\boldmath $\cal N$}}_{\rm prec}-{\mbox{\boldmath $\cal N$}}_{\rm warp}, \label{spin}\end{equation} where $I$ is the moment of inertia, ${\mbox{\boldmath $\cal N$}}_{\rm prec}$ and ${\mbox{\boldmath $\cal N$}}_{\rm warp}$ are the total warping and precessional torques acting on the disk \footnote{In (\ref{spin}) we have neglected possible magnetic torque [as in the Ghosh \& Lamb (1979a,b) picture] associated closed magnetic field lines that thread the disk and connect the star. Such magnetic torque can be included by modifying the first term on the right-hand-side of (\ref{spin}) to $\dot M\sqrt{GMr_m}\,f\,{\hat l}$, where $f$ is a dimensionless function which depends on the ratio $\omega/\Omega(r_m)$ [See, e.g., eq.~(35) of Lai (1998) and references therein].}. To calculate ${\mbox{\boldmath $\cal N$}}_{\rm prec}$ and ${\mbox{\boldmath $\cal N$}}_{\rm warp}$, we need to know the tilt angle $\beta$ as a function of $r$. In principle this can be obtained by solving the tilt equation (\ref{warp}) or (\ref{warp1}). But note that since the torques per unit area (eqs.~[\ref{torque3b}]-[\ref{torque3c}]) are steep functions of $r$, we can assume, as a first approximation, that $\beta$ is independent of $r$ near $r_m$. We then find \begin{equation} {\mbox{\boldmath $\cal N$}}_{\rm prec}=\int_{r_m}^\infty\!2\pi r {\bf N}_{\rm prec}\, dr =-{4\mu^2\over 3\pi r_m^3}\sin^2\!\theta \cos\beta\,({\hat\omega}\times{\hat l}), \end{equation} and \begin{equation} {\mbox{\boldmath $\cal N$}}_{\rm warp}=\int_{r_m}^\infty\!2\pi r {\bf N}_{\rm warp}\, dr =-{\zeta\,\mu^2\over 12 r_m^3}\cos^2\!\theta \sin 2\beta\,{\hat y}=-{\zeta\mu^2\over 6r_m^3}\cos^2\!\theta\cos\beta\,\left[ {\hat\omega}-({\hat\omega}\cdot{\hat l}){\hat l}\right]. \end{equation} Note that the ratio of the characteristic magnetic torque, ${\cal N}_{\rm mag}=\mu^2/r_m^3$, and the characteristic accretion torque, ${\cal N}_{\rm acc}=\dot M\sqrt{GMr_m}$, is \begin{equation} {{\cal N}_{\rm mag}\over{\cal N}_{\rm acc}}=\left({\mu^2\over\dot M\sqrt{GM}}\right) {1\over r_m^{7/2}}=\eta^{-7/2}, \end{equation} [see eq.~(\ref{alfven}) and Appendix B]. The spin evolution equation can then be written as \begin{eqnarray} {d\over dt}\left(I{\mbox{\boldmath $\omega$}}\right)&=&{\cal N}_{\rm acc} \left[1+{\zeta\cos^2\!\theta\over 6\eta^{7/2}}\sin^2\!\beta\right]\cos\beta \,\,{\hat\omega}\nonumber\\ &&+{\cal N}_{\rm acc}\left[{\zeta\cos^2\!\theta\over 6\eta^{7/2}}\cos^2\!\beta -1\right]\sin\beta\,\,{\hat\omega}_\perp\nonumber\\ &&-{\cal N}_{\rm acc}\left({4\over 3\pi\eta^{7/2}}\sin^2\!\theta\cos\beta \right)\,{\hat l}\times{\hat\omega}, \label{spin2}\end{eqnarray} where ${\hat\omega}_\perp$ is a unit vector perpendicular to ${\hat\omega}$ and lies in the plane spaned by ${\hat l}$ and ${\hat\omega}$ (see Fig.~1). The physical effects of the three terms on the right-hand-side of (\ref{spin2}) are evident: The first term is responsible to the spin-up or spin-down (depending on the sign of $\cos\beta$) of the star; the second term tends to align or misalign the spin axis with the disk axis\footnote{Since the timescale to change ${\hat\omega}$ is much longer than the timescale to change ${\hat l}$, the change of the tilt angle $\beta$ is determined by the dynamics of ${\hat l}$ rather than ${\hat\omega}$ (see eq.~[\ref{warp3}] and \S 4).}, and the third term induces precession of the star's spin axis ${\hat\omega}$ around ${\hat l}$. The timescales associated with these changes of ${\mbox{\boldmath $\omega$}}$ (spin-up/spin-down, alignment/misalignment and precession) are all of order \begin{equation} \tau_{\rm spin}={I\omega\over{\cal N}_{\rm acc}} ={I\omega \over \dot M\sqrt{GMr_m}}. \label{tspin}\end{equation} This spin-changing timescale is typically much longer than the timescale associated with changing the disk orientation (see \S 4 and \S6). Note that eq.~(\ref{tspin}) represents the instantaneous spin-up/spin-down timescale (i.e., at a given $\beta$). If the disk inclination $\beta$ wanders around $90^\circ$, we expect the secular spin-up of the star to be slower (see \S 6). \section{ASTROPHYSICAL APPLICATIONS} In this section we discuss/speculate several possible applications of our theory. While preliminary, they demonstrate the potential importance of the physical effects uncovered in previous sections. \subsection{Spin Evolution of Disk-Fed X-Ray Pulsars} Recent long-term, continuous monitoring of accreting X-ray pulsars with the BATSE instrument on the Compton Gamma Ray Observatory has revealed a number of puzzling behaviors of the spins of these accreting, magnetized neutron stars (see Bildsten et al.~1997 and references therein). Several well-measured disk-fed systems (e.g., Cen X-3, GX 1+4 and 4U 1626-67) display sudden transitions between episodes of steady spin-up and spin-down, with the absolute values of spin torques approximately equal (to within a factor of two). Of special interest is the observed anticorrelation between the torque and X-ray luminosity during the spin-down phase of GX 1+4 (i.e., the torque becomes more negative as the luminosity increases; Chakrabarty et al.~1997). These features are at odds with previous theoretical models, according to which the neutron star must be near spin-equilibrium in order to experience both spin-up and spin-down (e.g., Ghosh \& Lamb 1979a,b; see also Yi, Wheeler \& Vishniac 1997; Torkelsson 1998; Lovelace et al.~1998). It has been noted that the observational data can be nicely explained if the disk can somehow change its sense of rotation (Nelson et al.~1997) --- This poses a significant theoretical problem, since the disk formed in a Roche lobe overflow has a well-defined direction of rotation. It has been suggested (Van Kerkwijk et al.~1998) that the disk reversal may be caused by the radiation-driven warping instability (Pringle 1996), although the extent and timescale of the reversal remain unclear. (Recall that one expects the radiation-driven instability to operate only at large radii; Pringle 1996.) Also, since the radiation driven warping does not directly depends on the spin orientation, it is not clear why the disk prefers to wander around the perpendicular state ($\beta =90^\circ$). Here we suggest that the magnetically driven warping instability uncovered in this paper plays an important role in the determining the spin behaviors of accreting X-ray pulsars. The magnetosphere is located at \begin{equation} r_m=\eta\left({\mu^4\over GM\dot M^2}\right)^{1/7} =\left(3.4\times 10^8\,{\rm cm}\right)\,\eta\,\,\mu_{30}^{4/7}M_{1.4}^{-1/7} \dot M_{17}^{-2/7}, \label{alfven2}\end{equation} and we shall consider $\eta$ to be a constant of order unity (see Appendix B). Here $\mu_{30}$, $M_{1.4}$ and $\dot M_{17}$ are the neutron star's magnetic moment, mass and accretion rate in units of $10^{30}$~G$\,$cm$^3$, $1.4M_\odot$ and $10^{17}$~g~s$^{-1}$, respectively. Near the magnetosphere, the disk lies in the so-called ``middle'' (gas pressure and scattering dominated) region of the $\alpha$-disk solution (Shakura \& Sunyaev 1973; Novikov \& Thorne 1973). The surface density of the disk is \begin{equation} \Sigma=\left(1.7\times 10^3\,{\rm g~cm}^{-2}\right)\, \alpha^{-4/5}M_{1.4}^{1/5}\dot M_{17}^{3/5} r_8^{-3/5}{\cal J}^{3/5}, \end{equation} where $r_8=r/(10^8~{\rm cm})$, and the possible effect of a threaded magnetic field can be included in the definition of ${\cal J}$ (see Appendix A). The growth rate of the magnetic warping instability (eq.~[\ref{gammaw}]) is given by \begin{eqnarray} \Gamma_w &=&\left(0.035\,\,{\rm s}^{-1}\right) \left(\zeta\cos^2\theta\right)\,\alpha^{4/5} \mu_{30}^2\,M_{1.4}^{-7/10}\dot M_{17}^{-3/5}r_8^{-49/10}{\cal J}^{-3/5}\nonumber\\ &=&\left({1\over 5.3~{\rm day}}\right) \left(\zeta\cos^2\theta\right)\,\left({\alpha\over 0.01}\right)^{4/5} \mu_{30}^{-4/5}\dot M_{17}^{4/5}\left({r_m\over\eta r}\right)^{49/10} {\cal J}^{-3/5}. \label{gamma}\end{eqnarray} As discussed in \S 2 and \S 4, the magnetic torque tends to drive the inner region (near the magnetosphere) of the disk toward a perpendicular configuration ($\beta=90^\circ$) on a timescale of order $\Gamma_w^{-1}$. Indeed, the $\beta=\pi/2$ state represents an ``attractor''. In the idealized situation (each disk ring evolves independent of each other), the tilt of a nearly orthogonal disk evolves according to \begin{equation} {d\Delta\beta\over dt} =\Gamma_w\sin\beta\cos\beta\simeq -\Gamma_w\Delta\beta, \end{equation} where $\Delta\beta\equiv\beta-\pi/2$ (see eq.~[\ref{warp3}]). The star can then spin-up or spin-down, depending on the whether $\beta<90^\circ$ or $\beta>90^\circ$ [see eq.~(\ref{spin2})]: \begin{equation} I\dot\omega={\cal N}_{\rm acc} \left[1+{\zeta\cos^2\!\theta\over 6\eta^{7/2}}\sin^2\!\beta\right]\cos\beta. \end{equation} The characteristic spin-up/spin-down time scale (eq.~[\ref{tspin}]) is \begin{equation} \tau_{\rm spin}=\left(7.9\times 10^3\,{\rm yrs}\right)\eta^{-1/2} M_{1.4}^{-3/7}I_{45}\,\mu_{30}^{-2/7}\dot M_{17}^{-6/7} \left({1\,{\rm s}\over P_s}\right), \label{spinup}\end{equation} where $P_s$ is the spin period, and $I_{45}=I/(10^{45}\,{\rm g~cm}^2)$. In our picture, the observed sign switching of $\dot\omega$ in several X-ray pulsars is associated with the ``wandering'' of $\beta$ around the ``preferred'' value ($\beta=90^\circ$). Such ``wandering'' needs not be periodic: Consider a disk initially at $\beta=90^\circ$. Imagine that a perturbation in the accretion induces a negative (not necessarily small) $\Delta\beta=\beta-90^\circ$ --- This is achieved on a viscous timescale (which is of the same order of magnitude as $\Gamma_w^{-1}$ at the inner disk edge; see \S 4.1). The star spins up. The magnetic torque then drives $\beta$ toward $90^\circ$ on the timescale of $\Gamma_w^{-1}$, at which point another perturbation (which cannot be faster than $\Gamma_w^{-1}$) can induce another $\Delta\beta$, which can be either negative or positive, and the star will then continue to spin-up or switch to spin-down. We expect that the timescale of the switching between spin-up and spin-down is of order a few times $\Gamma_w^{-1}$ (evaluated at the inner disk boundary). For $\alpha\sim 0.1-0.01$ and $\dot M_{17}\sim 0.1$, $\mu_{30}\sim 1$ (typical of X-ray pulsars), eq.~(\ref{gamma}) gives $\Gamma_w^{-1}\sim 5-30$~days (for $\zeta\cos^2\!\theta=1$), comparable to what is observed in Cen X-3 ($P_s=4.8$~s) (Bildsten et al.~1997). For 4U 1626-67 ($P_s=7.6$) and GX 1+4 ($P_s=120$~s), the sign of $\dot\omega$ switchs once in $10-20$~yr. Since equation (\ref{gamma}) is uncertain and depends on many parameters, it is conceivable that such long switching time can be accommodated (e.g., with $\zeta\cos^2\!\theta=0.1$, $\alpha=0.01$, $\dot M_{17}=0.1$ and $\mu_{30}=10$, we find $\Gamma_w^{-1}\simeq 6$~years). Note that eq.~(\ref{spinup}) represents the ``instantaneous'' spin-up/spin-down time of the neutron star. Since the outer disk has a well-defined direction (presumably with $\beta<90^\circ$), the inner disk will unlikely spend equal amount of time in the prograde ($\beta<90^\circ$) phase and the retrograde ($\beta>90^\circ$) phase. Thus we expect that on a longer time scale (longer than the disk reversal time $\Gamma_w^{-1}$), the neutron star will experience secular spin-up. This expectation is borne out by observations, although it is difficult to predict the long-term spin-up rate. A full study or simulation of the nonlinear behavior of the inner disk tilt would be desirable to make more meaningful comparison with observations. But our discussion and estimate given above indicates that magnetically driven warping may be a crucial ingredient in explaining the spin behaviors of disk-fed X-ray pulsars. Other physical effects (such as radiation driven warping and propeller effect; Van Kerkwijk et al.~1998, Lovelace et al.~1998) may also play a role. \subsection{Weakly Magnetized Neutron Stars: Quasi-Periodic Oscillations} Rapid variability in low-mass X-ray binaries (LMXBs), containing weakly magnetized ($B\sim 10^7-10^{9}$~G) neutron stars, has been studied since the discovery of the so-called horizontal-branch oscillations (HBOs) in a subclass of LMXBs called Z sources (van der Klis et al.~1985; Hasinger \& van der Klis 1989). The HBOs are quasi-periodic oscillations (QPOs) (with the $Q$-value $\nu/\Delta\nu$ of order a few, and rms amplitude $\lo 10\%$) which manifest as broad Lorentzian peaks in the X-ray power spectra with centroid frequencies in the range of $15-60$~Hz which are positively correlated with the inferred mass accretion rate (see van der Klis 1995 for a review). For many years, the standard interpretation for the HBOs has been based on the magnetosphere beat-frequency model, first advocated by Alpar \& Shaham (1985) and Lamb et al.~(1985), in which the HBO is identified with the difference frequency between the Keplerian frequency at the magnetospheric boundary and the spin frequency of the neutron star (see Ghosh \& Lamb 1992 for a review). However, recent observations of kHz QPOs ($500-1200$~Hz) in at least eighteen LMXBs by the Rossi X-ray Timing Explorer (RXTE) have called into question this interpretation of HBOs (see van der Klis 1998a,b for review). The kHz QPOs (with $Q$ up to $100$ and rms amplitude up to $20\%$) often come in pairs, and both frequencies move up and down as a function of photon count rate, with the separation frequency roughly constant (The clear exceptions are Sco X-1 and 4U 1608-52; van der Klis et al.~1997, Mendez et al.~1998a). In most Z sources, the 15-60 Hz HBOs appear simultaneously with the kHz QPOs, while in several atoll sources (which are thought to have weaker magnetic fields and smaller accretion rates than the Z sources), broad peaks at $10-50$~Hz in the power spectra (similar to the HBOs in the Z sources) have also been detected at the same time when the kHz QPOs appear. While the origin of the kHz QPOs is uncertain, it is natural to associate the higher frequency QPO with the orbital motion at the inner edge (perhaps the magnetosphere boundary) of the accretion disk, and the lower-frequency QPO may result from the (perhaps imperfect) beat between the Kepler frequency and the neutron star spin frequency\footnote{See Stella \& Vietri (1999) for an alternative interpretation of the lower QPO peak which does not involve beating.} --- This beat frequency interpretation is supported by the observations of a third, nearly coherent QPO which occurs during X-ray bursts, at a frequency approximately equal to the frequency difference between the twin kHz peaks or twice that value. (An exception is 4U 1636-53, Mendez et al.~1998b.) Clearly, if this generic identification of kHz QPOs is correct, the beat between the $\sim 1$~KHz Keplerian frequency at the disk inner edge and the $\sim 300$~Hz spin frequency cannot produce the $10-60$~Hz low-frequency QPOs (LFQPOs: HBOs in the Z sources and similar features in the atoll sources), unless one postulates (Miller, Lamb \& Psaltis 1998) that the inner disk edge lies inside the magnetosphere and is unaffected by the magnetic field. Stella and Vietri (1998) suggested that the $10-60$~Hz LFQPOs are associated with Lense-Thirring precession of the inner accretion disk around the rotating neutron star. Assuming that the low-frequency QPO and the kHz QPOs are generated at the same radius in the disk, one obtains, to leading order, $\Omega_{\rm LT}=(2I\omega/Mc^2)\Omega^2$ (where $\Omega,~\Omega_{\rm LT}$ and $\omega$ are the orbital, Lense-Thirring, and spin angular frequencies, respectively). Thus the LFQPO frequency depends quadratically on the kHz QPO frequency. (The classical precession associated with spin-induced oblateness of the star, as well as relativistic effect beyond the Lense-Thirring formula, both can introduce small correction to this correlation; see Stella \& Vietri 1998, Morsink \& Stella 1999.) This is consistent with observations of several sources (e.g., the Z sources GX17+2, GX5-1, Sco X-1, and the atoll source 4U1728-34; see Ford \& van der Klis 1998, Psaltis et al.~1999 and references therein) \footnote{Observations indicate that the ratio $I/M$ required to fit the expected $\Omega_{\rm LT}-\Omega$ relation is a factor of $2-4$ larger than allowed by neutron star equation of state. This situation can be improved if the observed LFQPO frequency is the second harmonic of the fundamental precession frequency (See, e.g., Stella \& Vietri 1998, Morsink \& Stella 1999, Psaltis et al.~1999). In some Z sources, one requires that the observed HBO frequency is four times the fundamental precession frequency in order to produce reasonable $I/M$. Alternatively, the spin frequency is twice of what is inferred from the difference between the twin kHz QPOs --- This would make the beat frequency interpretation of the lower kHz peak invalid. It would be interesting to search for ``sub-harmonic'' feature of the HBOs in the power spectra (see Ford \& van der Klis 1998 for possible evidence of such a sub-harmonic feature in 4U 1728-34).}. For the Lense-Thirring interpretation of HBOs to be viable, the inner disk must be tilted with respect to the stellar spin axis. The Bardeen-Petterson effect tends to keep the inner region of the disk [inside $r_{\rm BP}$, typically at $(100-1000)GM/c^2$] co-planar with the star (Bardeen \& Petterson 1975). Radiation driven warping (Pringle 1996) is only effective at large disk radii. While global disk warping modes may exist with nonzero tilt near the inner disk boundary (Ipser 1996; Markovi\'c \& Lamb 1998), an external driving force is needed to excite them. Here we suggest that the magnetic warping torque provides a natural driver for the disk tilt near the inner accretion disk \footnote{Vietri and Stella (1998) suggested that if the accretion disk is inhomogeneous, diamagnetic blobs can be lifted above the equatorial plane through resonant interaction with the star's magnetic field near the the corotation radius. See Appendix C.}. For typical parameters of LMXBs, the magnetosphere is located at\footnote{Note that since $r_m$ is so close to the inner-most stable orbit, general relativistic effect tends move the inner disk edge to a radius larger than what is given in (\ref{alf2}); see Lai (1998). Here we shall neglect such complication.} \begin{equation} r_m=\left(18\,{\rm km}\right)\,\eta\,\mu_{26}^{4/7}M_{1.4}^{-1/7}\dot M_{17}^{-2/7} \label{alf2}\end{equation} where $\mu_{26}=\mu/(10^{26}\,{\rm G~cm}^3)$. Assuming that near the magnetosphere the disk is described by the ``inner region'' (radiation and scattering dominated) solution of $\alpha$-disk (Shakura \& Sunyaev 1973; Novikov \& Thorne 1973), we have for the disk surface density and half-thickness \begin{eqnarray} \Sigma &=&\left(105\,\,{\rm g~cm}^{-2}\right) \alpha^{-1}M_{1.4}^{-1/2}\dot M_{17}^{-1}r_6^{3/2}{\cal J}^{-1},\label{surf}\\ H&=& \left(1.1\,{\rm km}\right)\dot M_{17}\,{\cal J},\label{height} \end{eqnarray} where $r_6=r/(10^6\,{\rm cm})$. The growth rate of the magnetic warping instability (eq.~[\ref{gammaw}]) is then \begin{eqnarray} \Gamma_w &=&\left(555\,{\rm s}^{-1}\right)\left(\zeta\cos^2\!\theta\right) \alpha\,\mu_{26}^2\dot M_{17}\,r_6^{-7}{\cal J}\nonumber\\ &=&\left(1.0\,\,{\rm s}^{-1}\right)\left(\zeta\cos^2\!\theta\right) \left({\alpha\over 0.1}\right)M_{1.4}\,\mu_{26}^{-2}\dot M_{17}^3 \left({r_m\over\eta r}\right)^7{\cal J}. \end{eqnarray} The magnetic effect also results in a precessional torque on the disk. From eq.~(\ref{prec3}) and using (\ref{surf})-(\ref{height}), we find that the precession frequency associated with this magnetic torque is \begin{eqnarray} \nu_{\rm prec}&=&-\left(0.21\,{\rm Hz}\right) \left(\sin^2\!\theta\cos\beta\right)\left({\alpha\over 0.1}\right) M_{1.4}\,\mu_{26}^{-2}\dot M_{17}^3\left({r_m\over\eta r} \right)^7 D^{-1}{\cal J}\nonumber\\ &=&-\left(0.60\,{\rm Hz}\right)\left(\sin^2\!\theta\cos\beta\right) \left({\alpha\over 0.1}\right)M_{1.4}^{3/7}\mu_{26}^{-12/7} \dot M_{17}^{20/7}\eta^{1/2}\left({r_m\over\eta r}\right)^7 {\cal J}^{1/2}, \label{precc}\end{eqnarray} where in the second equality we have used $D=(2H/r_m)^{1/2}$ [see eq.~(\ref{Dfunc})]. This should be compared with the Lense-Thirring precession frequency \begin{eqnarray} \nu_{\rm LT}&=&\left(44.4\,{\rm Hz}\right)\,I_{45}\,r_6^{-3}\left({\nu_s \over 300\,{\rm Hz}}\right)\nonumber\\ &=&\left(8.1\,{\rm Hz}\right)M_{1.4}^{3/7}\,I_{45}\,\mu_{26}^{-12/7}\dot M_{17}^{6/7}\left({\nu_s\over 300\,{\rm Hz}}\right)\left({r_m\over \eta r}\right)^3, \end{eqnarray} where $\nu_s$ is the spin frequency. Note that for $\beta<90^\circ$ the magnetically driven precession is opposite to the stellar spin axis (thus the negative sign in eq.~[\ref{precc}]), while Lense-Thirring precession is in the same direction as the spin. For a given source, the magnitude of $|\nu_{\rm prec}|$ is typically smaller than $\nu_{\rm LT}$, but it becomes increasingly important with increasing $\dot M$ (since $\nu_{\rm LT}\propto\dot M^{6/7}$ while $\nu_{\rm prec}\propto\dot M^{20/7}$). This may explain the observed flattening of the correlation between the LFQPO frequency and kHz QPO frequency as the latter increases \footnote{Since the classical precession rate due to the oblateness of the star is negative (for $\cos\beta>0$) and scales as $r^{-7/2}$ (while $\nu_{\rm LT}\propto r^{-3}$) (Stella \& Vietri 1998, Morsink \& Stella 1999), it may also explain this observed flattening, provided that the spin frequency is much higher than inferred from the kHz QPOs and the burst QPOs. However, because of the similar power-law indices ($r^{-7/2}$ vs. $r^{-3}$), it is difficult to explain why the $\nu_{\rm LFQPO}\propto \nu_{\rm kHzQPO}^2$ scaling breaks down only at very high kHz QPO frequency (Psaltis et al.~1999).}. We note that although our qualitative conclusion that magnetic effect can induce tilt of the inner accretion disk and therefore facilitate its precession is robust, the analytical expressions given in this subsection only serve as an order-of-magnitude estimate which indicates the potential importance of the magnetically driven precession. More detailed calculations (including global mode analysis) are needed to make more meaningful comparison with observational data. In addition to the intrinsic uncertainties associated with the $\alpha$-disk model and the magnetosphere boundary layer, there are two complications that may prove important for LMXBs: (i) General relativistic effect can modify the inner radius of the disk so that it is not just determined by magnetic-plasma stress balance (Lai 1998); (ii) The magnetic field is not expected to be dipolar. This comes about either because of the multipole fields from the central star, or, even if the intrinsic stellar field is dipolar, the (partially diamagnetic) disk can enhance the field in the boundary layer (Lai, Lovelace \& Wasserman 1999). \subsection{Super-Orbital Periods in X-ray Binaries} Here and in \S 6.4 we speculate upon two additional applications of our theory. Many X-ray binaries are known to exhibit long-term cycles in their X-ray or optical luminosities (see Priedhorsky \& Holt 1987 and references therein). Of particular interest is the well-established ``third'' period (longer than the orbital period) in Her X-1 (35 d) (Tananbaum et al.~1972), LMC X-4 (30.5 d) (Lang et al.~1981) and SS433 (164 d) (Margon 1984). It is generally thought that these super-orbital periods result from the precession of a tilted accretion disk. For Her X-1, Katz (1973) proposed that the precession was forced by the torque from the companion star, but left unexplained the origin of the disk's misalignment with respect to the orbital plane. Recently it has been suggested that the radiation-driven warping instability (Pringle 1996) is responsible for producing warped, precessing disks in many X-ray binaries which exhibit long-term cycles (Maloney \& Begelman 1997; Wijers \& Pringle 1998). It is likely that magnetic torque plays a role in driving the warping of the inner disk. The observed systematic variation of the X-ray pulse profile of Her X-1 requires the inner edge of the disk to be significantly warped (Sheffer et al.~1992; Deeter et al.~1998 and references therein). In addition, the magnetically driven precession frequency is \begin{equation} \nu_{\rm prec}=-\left({1\over 26\,{\rm day}}\right) \left(\cos\beta\sin^2\!\theta\right)\,\left({\alpha\over 0.01}\right)^{4/5} \mu_{30}^{-4/5}\dot M_{17}^{4/5}\left({r_m\over\eta r}\right)^{49/10} {\cal J}^{-3/5}D^{-1}, \label{precf}\end{equation} where $r_m$ is the magnetosphere radius [eq.~(\ref{alfven2})], and we have adopted the ``middle region'' solution of the $\alpha$-disk (near the inner edge of the disk, $D\sim 0.2$). This is comparable to the observed super-orbital periods. Moreover, the magnetically driven precession is retrograde with respect to the direction of rotation of the disk, in agreement with observations. Of course, the precession rate is a function of $r$ and $\beta$, so a modal analysis is needed to determine the global precession period. \subsection{T Tauri Stars} It is well established that classical T Tauri stars (CTTS) have circumstellar disks; Evidence also exists for magnetospheric accretion induced by the stellar magnetic field (e.g., Hartmann 1998). For typical parameters ($M\simeq 0.5\,M_\odot$, $R\simeq 2R_\odot$, $\dot M=10^{-9}-10^{-7}\,M_\odot$~yr$^{-1}$, and surface field $B_\star\sim 1$~kG), the magnetosphere is located at a few stellar radii. Using $\mu=B_\star R^3$, we find that the growth time for the warping instability at the inner region of the disk is given by [see eq.~(\ref{gammaw})] \begin{equation} \Gamma_w^{-1}=(6.5\,{\rm days})\left({1\,{\rm kG}\over B_\star}\right)^2\! \left({2R_\odot\over R}\right)^6\!\left({M\over 0.5M_\odot}\right)^{1/2}\!\! \left({r\over 8R_\odot}\right)^{11/2}\!\! \left({\Sigma\over 1\,{\rm g\,cm}^{-2}}\right)\! \left(\zeta\cos^2\theta\right)^{-1}. \end{equation} The precession period of the tilted disk (see eq.~[\ref{prec3}]) is of the same order of magnitude as $\Gamma_w^{-1}$. The surface density $\Sigma$ is unknown, but reasonable estimates give $\Sigma\sim 1-100$~g~cm$^{-2}$ (Hartmann 1998). Therefore $\Gamma_w^{-1}$ ranges from days to years, much shorter than the lifetime of T Tauris stars ($\sim 10^6$~years). It has been observed that the photometric periods ($5-10$~days) of CTTS vary (by as much as $30\%$) on a timescale of weeks (Bouvier et al.~1995). The origin of this variability is unknown. If we interprate the photometric period as the orbital period at the magnetosphere boundary (Bouvier et al.~1995), then we may understand the period variation in the context of warped, precessing disks: As the inner disk warps and precesses, it experiences different stellar magnetic field, and thus the inner disk radius varies. It is also of interest to consider the effect of magnetically driven warping on the rotation of T Tauri stars. The projected rotation velocity of CTTS with masses $M\lo 1\,M_\odot$ is about $20$~km~s$^{-1}$, only $10\%$ of the breakup speed (e.g., Bertout 1989). This is at odds with the expectation that T Tauri stars are formed by the gravitational collapse of rotating molecular cores and the presence of disks surrounding the stars. Theories which explain the slow rotations generally invoke the interaction between the disk and the stellar magnetic field of a few kG (e.g., K\"onigl 1991; Cameron \& Campbell 1993; Shu et al.~1994; Yi 1995; Armitage \& Clarke 1996). Since the growth time for disk warping is short, we may expect the inner disk of T Tauri stars to wander around the ``preferred'' perpendicular state. The star therefore experiences both spin-up and spin-down during its evolution, analogous to the behavior of X-ray pulsars (see \S 6.1). The net, secular spin-up rate is expected to be much smaller than that based on the canonical spin-up torque ($\sim \dot M\sqrt{GMr_m}$). \section{CONCLUSION} In this paper we have identified a new magnetically driven warping instability which occurs in the inner accretion disk of a magnetized star (\S 2 and \S 4). Despite the uncertainties in our understanding of the magnetosphere--disk interactions (particularly the global magnetic field structure), the existence of the instability seems robust, and requires that some vertical field lines (either from the star or intrinsic to the disk) thread the disk and get twisted by the disk rotation. The general consequence of the instability is that the normal vector of the inner disk (near the magnetosphere) will be tilted with respect to the stellar spin axis. We have also shown that the disk can be driven magnetically into precession around the spin axis due to the interaction between the screening surface current and the stellar magnetic field (\S 2). In addition, certain regions of the disk are subjected to resonant magnetic forces which may affect the structure and dynamics of the disk (\S 3). These magnetic effects on accretion disks have largely been overlooked in previous studies of accretion onto magnetic stars \footnote{We note that the effects studied in this paper are quite different from the situation considered by Agapitou et al.~(1997), who studied the bending instability in the inner (sub-Keplerian) disk which corotates with the star (with the magnetic axis aligned with the spin axis). Such a disk may or may not exist (see Spruit \& Taam 1990).}. We have applied our theory to several different types of astrophysical systems, including X-ray pulsars, low-mass X-ray binaries, and T Tauri stars (\S 6). These applications should be considered preliminary, but they indicate that the magnetically driven disk warping and precession can potentially play an important role in determining the observational behaviors of these systems. Of particular interest is that the magnetically warped inner disk may provide a natural explanation for the longstanding puzzle of torque reversal as observed in a number of X-ray pulsars. Also, the tilted disk may be responsible for the rich phenomenology of time variability (such as QPOs) observed in weakly magnetized accreting neutron stars in LMXBs. Much work is needed to understand better the effects studied in this paper and their observational manifestations in different astrophysical systems. In our analysis, we have intentionally avoided (or bypassed), by using parametrized models, the uncertainties associated with magnetosphere -- disk interactions (see, e.g., Appendix A,B), but clearly the study of disk warping and precession based on more specific models (with or without outflows) would be useful. The role of intrinsic disk field needs to be examined further (see \S 2.4). There remains uncertainty in the description of nonlinear warped disks, and a full numerical simulation of the nonlinear development of the warping instability would be valuable. In the case of low-mass X-ray binaries, the effects of complex magnetic field topology (other than dipole) and general relativity should be included to access the QPO phenomenology (see \S 6.2). More detailed comparison with observational data will be useful. The role of magnetically driven resonances need to be studied further to determine whether they will produce any observable features. We hope to address some of these issues in the future. \acknowledgments I thank Richard Lovelace, Phil Maloney, Dimitrios Psaltis, Marten van Kerkwijk and Ethan Vishniac for useful discussion/comment. I also thank Brad Hansen for informing me of some important references. I became puzzled by the spin-up/spin-down behavior of X-ray pulsars several years ago at Caltech through discussions with Lars Bildsten, Deepto Chakrabarty, Rob Nelson and Tom Prince. I acknowledge the support from a NASA ATP grant and a research fellowship from the Alfred P. Sloan foundation, as well as support from Cornell University.
\section*{Introduction} The configuration space $\Conf(X,n)$ of $n$ distinct labeled points in a topological space~$X$ is the complement in the Cartesian product~$X^n$ of the union of the large diagonals $`D^{ij}=\{(x_1,\dots,x_n)\thinspace|\thinspace x_i=x_j\}$. Pioneering studies of these spaces by Fadell, Neuwirth, Arnold and Cohen \cite{Arnold,FCohen,Fadell,FadellNeuwirth} evolved into a still active area of algebraic topology; Totaro opens his paper with a brief review~\cite{Totaro}. Somewhat later, a compactification of $\Conf(\C,n)$ modulo affine automorphisms, known as the Grothendieck--Knudsen moduli space of stable $n$-pointed curves of genus~0, rose to prominence in modern algebraic geometry \cite{Deligne:resume,Kapranov:Veronese,Keel,Knudsen}. Then Fulton and MacPherson\ devised a powerful construction that works for any nonsingular algebraic variety and produces a compactification $X[n]$ of $\Conf(X,n)$ with a remarkable combination of properties~\cite{FM}: \begin{list}{$\triangleright$}{} \item $X[n]$ is nonsingular. \item $X[n]$ naturally comes equipped with a proper map onto $X^n$. \item $X[n]$ is symmetric: it carries an action of the symmetric group~$\Sym_n$ by permuting the labels. \item The complement $D=X[n]\smallsetminus \Conf(X,n)$ is a normal crossing divisor. \item The combinatorial structure of~$D$ and of the resulting stratification of~$X[n]$ is explicitly described: the components of~$D$ correspond to the subsets of $[n]=\{1,\dots,n\}$ with at least~2 elements; their intersections, the strata, correspond to nested collections of such subsets, and the latter are just a reincarnation of rooted trees with~$n$ marked leaves. \item Degenerate configurations have simple geometric descriptions. \end{list} Further results of Fulton and MacPherson\ include: a functorial description of~$X[n]$, used to prove many of its properties listed above; a fact that all isotropy subgroups of~$\Sym_n$ acting on~$X[n]$ are solvable; some intersection theory, namely, a presentation of the intersection rings of $X[n]$ and of its strata, and, as an application, a computation of the rational cohomology ring of $\Conf(X,n)$ for~$X$ a smooth compact complex variety. About the same time, constructions related to the Fulton--MacPherson compactification appeared, all motivated by, and suited to, some problems of mathematical physics: for real manifolds \cite{AxelrodSinger,Kontsevich:Feynman}; for complex curves~\cite{BeilinsonGinzburg}, with later extension to higher dimensions~\cite{Ginzburg}. The compactifications $X[n]$ are defined inductively, with the step from $X[n]$ to $X[n+1]$ performed by a sequence of blowups $$ \xymatrix @+3pt { X[n+1]= Y_{n}\ar[r]^<(0.3){{}\alpha_{n-1}} & Y_{n-1}\ar[r]^<(0.35){{}\alpha_{n-2}} & {}\dots\ar[r]^{{}\alpha_1} & Y_1 \ar[r]^<(0.2){{}\alpha_0} & Y_0=X[n]\times X, } $$ where the center of the blowup~$`a_k$ is a disjoint union of subvarieties in~$Y_k$ corresponding in a specified way to the subsets of~$[n]$ of cardinality $n-k$. Thus, the symmetry of $\Sym_{n+1}$ is not present at the intermediate stages. An alternative, and completely symmetric, description of~$X[n]$ as the closure of $\Conf(X,n)$ in a product of blowups does not provide much insight into the structure of~$X[n]$, so the inductive sequence of blowups is essential for that. Fulton and MacPherson remark: \begin{quote} {\small It would be interesting to see if other sequences of blowups give compactifications that are symmetric, and whose points have explicit and concise descriptions \cite[bottom of p.\! 196]{FM}. } \end{quote} An example of such a compactification, for any nonsingular algebraic variety~$X$, is studied in the present paper. I~denote it by~$\Xn$ and call it a {\em polydiagonal compactification}, because the blowup loci are not only the diagonals of~$X^n$, but also their {\em intersections}. The idea is very simple: one who tries to blow up all diagonals of the same dimension simultaneously is forced to blow up all their intersections prior to that, and this prescribes the sequence. Following Fulton and MacPherson's terminology, $\Xn$ is a compactification even though it is only compact when~$X$ itself is compact. In general, it is equipped with a canonical proper map onto~$X^n$. This construction applies also to real manifolds, with real blowups replacing algebraic blowups. The compactification is then a manifold with corners, and the results about the strata presented here can be rephrased to describe the combinatorics of its boundary. The construction of~$\Xn$ is in some respects similar to that of~$X[n]$, with one important difference: the former is completely symmetric {\em at each stage}. This reduces logical complexity of the construction even though it involves (considerably) more blowups. From this last fact stems another feature of~$\Xn$: it distinguishes some collisions that are treated as equal by Fulton and MacPherson. There is a surjection $\vartheta_n \colon \Xn\to X[n]$ that essentially retreats from making these distinctions, and it is completely symmetric as well. Regardless of~$X$, this map, derived from a description of~$\Xn$ as the closure of $\Conf(X,n)$ in a product of blowups, is an isomorphism for $n\<3$ only, and an iterated blowup otherwise. The fibers of~$\vartheta_n$ have purely com\-bi\-na\-to\-ri\-al nature and do not depend even on the dimension of~$X$; their detailed description will appear in a separate paper~\cite{U2}. Geometrically, the limiting configurations in the Fulton--MacPherson compactification are viewed in terms of tree-like successions of {\em screens}, each of which is a tangent space to~$X$ with several labeled points in it, considered modulo translations and dilations. In a similar visualization for points in~$\Xn$, labels of a new kind, necessary because~$\Xn$ has \inquotes{more} points than~$X[n]$, augment the screens. This rests on a study of the strata: they are bundles over $X\pd<r>$, $r < n$, with fibers decomposable into products of certain projective varieties. Named {\em bricks}, they form a family indexed by integer partitions that includes, for example, permutahedral varieties. The latter in fact show up in each brick as constituents that account for those new labels. As for the combinatorics underlying~$\Xn$, here the place of subsets of~$[n]$, nested collections of such subsets, and plain rooted trees is taken by partitions of the set~$[n]$, chains of such partitions, and rooted trees whose vertices are assigned integer numbers, called {\em levels}. With these changes, the natural stratification of~$\Xn$ is quite similar to that of~$X[n]$; moreover, $\vartheta_n$~is a strata-compatible map corresponding to the forgetful map from leveled trees to usual rooted trees. Analogues for~$\Xn$ of most results of Fulton and MacPherson\ follow purely geometrically. Since the proofs do not require a functorial description of the space, it is omitted. The action of the symmetric group~$\Sym_n$ on~$X^n$ by permuting the labels has fixed points. Fulton and MacPherson showed that the isotropy subgroups of the label permutation action of~$\Sym_n$ on~$X[n]$ are solvable \cite[Theorem~5]{FM}. It turns out that in characteristic~0 the similar action of~$\Sym_n$ on~$\Xn$ has only abelian isotropy subgroups; thus, singularities of $\Xn\!/\Sym_n$ can in principle be resolved by toric methods \cite{Ash+,JLB:eventails,Kempf+,Oda}. The resulting space will provide an explicit desingularization of the symmetric product $X^n\!/\Sym_n$, as well as a smooth compactification of $\B(X,n)=\Conf(X,n)/\Sym_n$, the configuration space of~$n$ {\em unlabeled} points in~$X$. De Concini and Procesi developed a general approach to compactifying complements of linear subspace arrangements by iterated blowups \cite{DeCP}. For each arrangement, it yields a family of {\em wonderful}\/ blowups with minimal and maximal elements. Although they work with linear subspaces, their technique is local and can be applied to $X^n \smallsetminus \Conf(X,n)$ for any smooth variety~$X$; in this case, the Fulton--MacPherson compactification is the minimal one, while the polydiagonal compactification is the maximal one. Along the lines of De Concini, MacPherson and Procesi~\cite{MacPP}, Yi Hu has extended many results presented here in Sections \ref{sec-construct},~\ref{sec-hodge} and~\ref{sec-closure} to blowups of arrangements of smooth subvarieties and then recovered Kirwan's partial desingularization of geometric invariant theory quotients~\cite{Hu,Kirwan}. In addition, Hu computed the intersection rings in that general context of arrangements. In the case of $\Xn$ these rings may be used to build a differential graded algebra model of $\Conf(X,n)$ for~$X$ a compact complex algebraic manifold, as Fulton and MacPherson\ did. After that, Kriz streamlined their differential graded algebra, while Totaro extracted a presentation of the cohomology ring of the configuration space from the Leray spectral sequence of its embedding into its \inquotes{naive}\ compactification~$X^n$~\cite{Kriz,Totaro}. \subsection*{Historical note} (Communicated by W.~Fulton.) Fulton and MacPherson\ sought to build the space whose points would be described by screens; early attempts led them to consider the spaces denoted here by $X\pd<4>$ and~$X\pd<5>$, and to identify what to blow down to create the desired $X[4]$ and~$X[5]$. Seeing that as $n$ grows, the blowdown description quickly becomes unwieldy, they chose not to pursue this in general and finally settled on a nonsymmetric procedure. D.~Thurston pointed out a symmetric construction of $X[n]$ and used its real analogue in his work on knot invariants~\cite{Thurston}. \subsection*{Standing assumptions} Throughout the paper, $X$ is a smooth irreducible $m$-dimensional ($m > 0$) algebraic variety over some field~$\Bbbk$, and $n$ is the number of labeled points in~$X$. The section on Hodge polynomials applies only to complex varieties, and that on the symmetric group action, only to the characteristic~0 case. \subsection*{Outline of the paper} The first section is informal and serves to introduce the basic ideas of the polydiagonal compactification on the simplest example. A combinatorial interlude of Section~\ref{sec-combinat} is followed by a discussion of polyscreens and colored screens that represent points in~$\Xn$. Formally stated and proved results begin in Section~\ref{sec-construct} that contains: construction of $\Xn$ by a symmetric sequence of blowups, a description of the combinatorics of the complement $\Xn\smallsetminus\Conf(X,n)$ as a divisor with normal crossings and of the ensuing stratification of~$\Xn$, and a recurrent formula for the number of the strata. If~$X$ is a complex variety, the blowup construction translates into a formula for the (virtual) Hodge polynomial\ $e(\Xn)$ in terms of~$e(X)$ derived in the next section. In Section~\ref{sec-closure}, a consideration of~$\Xn$ as the closure of $\Conf(X,n)$ in a product of blowups implies a surjection $\vartheta_n \colon \Xn\to X[n]$, written then as an iterated blowup. Technical analysis of the strata of~$\Xn$ occupies Section~\ref{sec-strata}, and the last section deals with the isotropy subgroups of~$\Sym_n$ acting on~$\Xn$. \subsection*{Acknowledgements} In many ways, I~am indebted to Jean--Luc Brylinski, my Ph.D.~advisor. I~am most grateful to William Fulton and Jim Stasheff for sharing their advice and for many valuable comments and insights. The referee's suggestions helped me improve exposition. I~would also like to thank Dmitry Tamarkin for useful discussions. \section{Small numbers of colliding points} \label{sec-x4} The purpose of this section is to introduce the main ideas of the paper by looking at the case of 4 points---% the smallest integer~$n$ for which $\Xn$ is different from $X[n]$ is~4. To begin with, consider an example of two collisions of four points in $X=`C^2$. The corresponding two limiting configurations arising in the approach of Fulton and MacPherson\ coincide; however, the polydiagonal compactification will distinguish them. Take four points labeled by 1 through 4 and make them collide as $t\to 0$ in the following way: \begin{list}{$\diamond$}{} \item the distance between 1 and 2 is $O(t^3)$, \item the distance between 3 and 4 is $O(t^2)$, \item the distance between the two pairs (12) and (34) is $O(t)$. \end{list} Then do the same thing, except for a small exchange: \begin{list}{$\diamond$}{} \item the distance between 1 and 2 is $O(t^2)$, \item the distance between 3 and 4 is $O(t^3)$, \end{list} and call the two limiting points $\x_1$ and $\x_2$. Both limiting points lie in the same stratum of~$X[4]$, the intersection of three divisors $D(1234)$, $D(12)$, and $D(34)$. The dimension of this stratum is~5; the dimension of its fiber over a point in the small diagonal $`D\subset X^4$ is~3. The three parameters record the \inquotes{directions}\ of collisions encoded by the middle tree in Figure~\ref{fig-not-same}. Specifying these directions for the two approach curves, that is, vectors hidden behind the symbol~$O$, one can arrange that $\x_1=\x_2$ in $X[4]$. \begin{figure}[b] \label{fig-not-same} $$ \xy (-30,0)*{\ }; (94,2)*{\parbox[b]{1.6in}{\caption{}}}; (-6,-6.4)*{{}^1}, (-2,-6.4)*{{}^2}, (2,-6.4) *{{}^3}, (6,-6.4) *{{}^4}, @={(-6,-4),(-4,0),(-2,-4),(-4,0),(0,8),(2,4),(2,-4),(2,4),(6,-4)}, s4="prev" @@{;"prev";**@{-}="prev"}; (24,-6.4)*{{}^1}, (28,-6.4)*{{}^2}, (32,-6.4) *{{}^3}, (36,-6.4) *{{}^4}, @i; @={(24,-4),(27,2),(28,-4),(27,2),(30,8),(33,2),(32,-4),(33,2),(36,-4)}, s4="prev" @@{;"prev";**@{-}="prev"}; (54,-6.4)*{{}^1}, (58,-6.4)*{{}^2}, (62,-6.4) *{{}^3}, (66,-6.4) *{{}^4}, @i; @={(54,-4),(58,4),(58,-4),(58,4),(60,8),(64,0),(62,-4),(64,0),(66,-4)}, s4="prev" @@{;"prev";**@{-}="prev"}; (22,2) \ar @{~>} (38,2); (52,2) \ar @{~>} (8,2) \endxy $$ \end{figure} These approach curves actually belong to a whole family~$\mathcal F$ of curves in $\Conf(X,4)$ whose limits in $X[4]$ may coincide. Indeed, consider the diagonals $`D^{12}$ and $`D^{34}$ in $X^4$, and their intersection $`D^{12|34}$. Both curves approach this intersection, but the first one does it while having a 3rd degree osculation to $`D^{12}$, and the second one does the same with $`D^{34}$. The projectivized normal space $\P(T_p X^4/T_p `D^{12|34})$ parametrizes the family, and the two curves above correspond to normal directions going along $`D^{12}$ and $`D^{34}$ respectively. This suggests looking into a possibility of involving blowups of subvarieties like $`D^{12} \cap `D^{34}$, if the objective is to obtain a compactification that would distinguish from one another collisions produced by curves in such families. The space that achieves this results from implementing a simple idea of blowing up \inquotes{from the bottom to the top}. Although the dominant feature of the general case first comes to light when $n=4$, it may be useful to begin with the cases of two and three colliding points. Assume that $\dim X > 1$. There is no ambiguity about the case of $n=2$ points: the compactification is the blowup of the diagonal in~$X^2$. If~$n=3$, blowing up the small diagonal $`D \subset X^3$ creates disjoint proper transforms of $`D^{12}$, $`D^{13}$ and~$`D^{23}$ that can then be blown up in any order. The resulting compactification coincides with~$X[3]$. For~$n>3$, however, this strategy will not work, and some additional blowups are needed \cite[bottom of p.~196]{FM}, but what they are Fulton and MacPherson\ do not specify. The left graph in Figure~3 shows the diagonals in~$X^4$, including the space itself, as vertices, and (nonrefinable) inclusions of the diagonals into each other as edges. As before, blow up the small diagonal first, then blow up the (disjoint) proper transforms of the four larger diagonals, like $`D^{123}$. Now try to blow up the next level below them simultaneously. It does not work: these six largest diagonals have not been made disjoint. How can this be fixed? \begin{figure}[t] \label{fig-six-lines} \def\kern-1pt{\kern-1pt} \xy (-40,10)*{\ }; (0,-8)*{\ }; (46,10)*{\parbox[b]{1.6in}{\caption{}}}; (0,-6); (0,18)**\dir{-}; (-14,0); (14,0)**\dir{-}; (11.4,-5.1); (-3.6,17.4)**\dir{-}; (-11.4,-5.1); (3.6,17.4)**\dir{-}; (14,-3); (-10,9)**\dir{-}; (-14,-3); (10,9)**\dir{-}; (0,0)*{\mydot}; (0,4)*{\mydot}; (8,0)*{\mydot}; (4,6)*{\mydot}; (0,12)*{\mydot}; (-4,6)*{\mydot}; (-8,0)*{\mydot}; (3.9,15.1)*{\scriptstyle 1\!2}; (9,-4)*{\scriptstyle 1\!3}; (-12.6,1.2)*{\scriptstyle 1\!4}; (-1.4,-4.2)*{\scriptstyle 2\03}; (9.5,7.2)*{\scriptstyle 2\04}; (-8.2,9.7)*{\scriptstyle 3\04}; \endxy \end{figure} \begin{figure}[b] \label{fig-x4} \def\kern-1pt{\kern-1pt} \hskip3mm \xy (-28,14)*{\ }; (60,-18)*{ \mycaption{6in}{Diagonals (left) and polydiagonals (right) in~$X^4$}}; (0,12)="1234", (-18,4)="123", (-6,4)="124", (6,4)="134", (18,4)="234", (-15,-4)="12", (-9,-4)="13", (-3,-4)="23", (3,-4)="14", (9,-4)="24", (15,-4)="34", (0,-12)="no", @={"1234","123","12","124","24","234","34","134","1234", "234","23","123","13","134","14","124","1234"}, s0="prev" @@{;"prev";**@{-}="prev"}, @i @={"12","13","23","14","24","34"}, "no"; @@{**@{-}}, @i @={"1234","123","124","134","234","12","13","14","23","24","34","no"}, @@{*{\mydot}}, "1234"+(0,1)*{{}^{1\02\03\04}}, "123"+(0,1)*{{}^{1\02\03}}, "124"+(-.5,1)*{{}^{1\02\04}}, "134"+(.5,1)*{{}^{1\03\04}}, "234"+(0,1)*{{}^{2\03\04}}, "12"+(-2,-1)*{{}^{1\02}}, "13"+(-2,-1)*{{}^{1\03}}, "23"+(-2,-1)*{{}^{2\03}}, "14"+(2,-1)*{{}^{1\04}}, "24"+(2,-1)*{{}^{2\04}}, "34"+(2,-1)*{{}^{3\04}}, (70,0)="m", "m"+(0,12)="1234", "m"+(-6,5)="12-34", "m"+(0,5)="13-24", "m"+(6,5)="14-23", "m"+(-18,4)="123", "m"+(-12,4)="124", "m"+(12,4)="134", "m"+(18,4)="234", "m"+(-15,-4)="12", "m"+(-9,-4)="13", "m"+(-3,-4)="23", "m"+(3,-4)="14", "m"+(9,-4)="24", "m"+(15,-4)="34", "m"+(0,-12)="no", @i @={"1234","123","12","124","24","234","34","134","1234", "234","23","123","13","134","14","124","1234"}, s0="prev" @@{;"prev";**@{-}="prev"}, @i @={"12-34","13-24","14-23"}, "1234"; @@{**@{-}}, @i @={"no","12","12-34","34","no","23","14-23","14","no", "13","13-24","24","no"}, s0="prev" @@{;"prev";**@{-}="prev"}, @i @={"1234","123","124","134","234","12-34","13-24","14-23", "12","13","14","23","24","34","no"}, @@{*{\mydot}}, "1234"+(0,1)*{{}^{1\02\03\04}}, "123"+(0,1)*{{}^{1\02\03}}, "124"+(-.5,1)*{{}^{1\02\04}}, "134"+(.5,1)*{{}^{1\03\04}}, "234"+(0,1)*{{}^{2\03\04}}, "12"+(-2,-1)*{{}^{1\02}}, "13"+(-2,-1)*{{}^{1\03}}, "23"+(-2,-1)*{{}^{2\03}}, "14"+(2,-1)*{{}^{1\04}}, "24"+(2,-1)*{{}^{2\04}}, "34"+(2,-1)*{{}^{3\04}}, \endxy \end{figure} The six lines intersecting at seven points depicted in Figure~2 are the images of the large diagonals of~$`R^4$ in the real projective plane $\P(`R^4/`D)$, where~$`D$ is the small diagonal. Four of the points correspond to diagonals like $`D^{123}$, and the other three, where the intersections are normal, represent additional loci that need to be blown up to make the large diagonals disjoint. The second graph in Figure~3 is obtained from the first one by adding these three intersections $`D^{12} \cap `D^{34}$, $`D^{13} \cap `D^{24}$ and $`D^{14} \cap `D^{23}$. All seven vertices in the second row correspond to subvarieties pairwise disjoint after the blowup of the small diagonal $`D\subset X^4$, so they can be blown up simultaneously, and---crucially---after that the subvarieties from the row just below become disjoint and can be blown up simultaneously. This gives a compactification $X\pd<4>$ of~$\Conf(X,4)$. \begin{figure}[t] \label{fig-FM-12-34} \xy (30,-7)*{\mycaption{3in}{A point in $X[4]$}}; (-5,29); (18,29)**\crv{(-15,18)&(4,20)&(10,26)&(30,18)&(26,26)}; (-2,25)*{X}; (12,26)*{\mydot}; @={(8,12),(8,22),(18,22),(18,12)}; s0="prev" @@{;"prev";**@{-}="prev"}; (8,22); (11.5,26.1)**\dir{.}; (18,22); (12.3,26.2)**\dir{.}; (14.5,15)*{\mydot}; (11.5,19)*{\mydot}; @i; @={(0,0),(0,10),(10,10),(10,0)}; s0="prev" @@{;"prev";**@{-}="prev"}; (0,10); (11.2,19.4)**\dir{.}; (10,0); (11.9,19)**\dir{.}; (2,6)*{\mydot} -(0,1.3)*{\scriptscriptstyle 1}; (8,4)*{\mydot} -(0,1.3)*{\scriptscriptstyle 2}; @i; @={(16,0),(16,10),(26,10),(26,0)}; s0="prev" @@{;"prev";**@{-}="prev"}; (16,0); (14.1,15)**\dir{.}; (26,10); (14.7,15.4)**\dir{.}; (18,3)*{\mydot} -(0,1.3)*{\scriptscriptstyle 3}; (24,7)*{\mydot} -(0,1.3)*{\scriptscriptstyle 4}; (70,2)="m" +(33,-8.5)*{\mycaption{2in}{Dilation}}; @i; @={"m"+(0,0),"m"+(16,0),"m"+(16,16),"m"+(0,16)}; s0="prev" @@{;"prev";**@{-}="prev"}; "m"+(5,6)*{\mydot} +(1,-1)*{\scriptscriptstyle 3}; "m"+(11,10)*{\mydot} -(1,-1)*{\scriptscriptstyle 4}; "m"+(11.7,-1.8)*{\scriptstyle\alpha_{34}=2}; "m"+(30,0)="m"; @i; @={"m"+(0,0),"m"+(16,0),"m"+(16,16),"m"+(0,16)}; s0="prev" @@{;"prev";**@{-}="prev"}; "m"+(2,4)*{\mydot} +(1,-1)*{\scriptscriptstyle 3}; "m"+(14,12)*{\mydot} -(1,-1)*{\scriptscriptstyle 4}; "m"+(12,-1.8)*{\scriptstyle\alpha_{34}=1}; "m"+(-7,8)*{=}; \endxy \end{figure} The construction of $X\pd<4>$ involves three more blowups than that of $X[4]$, so the complement of $\Conf(X,4)$ in $X\pd<4>$ has three additional components $D^{12|34}$, $D^{13|24}$ and $D^{14|23}$. Collisions belonging to the family~$\mathcal F$ discussed above result in points in $Z = D^{1234}\cap D^{12|34}$. To accommodate these, as well as more complicated degenerations of the same nature that appear for $n>4$, two new features are added to Fulton--MacPherson screens: the screens are grouped into \newterm{levels}, and the group on each level bears a new parameter living in a projective space. Figure~4 illustrates the screen description of the limiting points in $X[4]$ of the family~$\mathcal F$: its macroscopic part is a single point in~$X$ and its microscopic part consists of three screens, one for each of the subsets $1234$, $12$~and~$34$ of $\{1,2,3,4\}$. A~screen is a tangent space $T_p X$ with a configuration of points in it, considered up to dilations and translations. In particular, the last two screens, $S_{12}$ and $S_{34}$, are completely independent of each other. \begin{figure}[b] \label{fig-new-level} \xy (30,-12)*{\mycaption{3in}{A degeneration in $X\pd<4>$}}; (-5,29); (18,29)**\crv{(-15,18)&(4,20)&(10,26)&(30,18)&(26,26)}; (-2,25)*{X}; (12,26)*{\mydot}; @={(8,12),(8,22),(18,22),(18,12)}; s0="prev" @@{;"prev";**@{-}="prev"}; (8,22); (11.5,26.1)**\dir{.}; (18,22); (12.3,26.2)**\dir{.}; (14.5,15)*{\mydot}; (11.5,19)*{\mydot}; @i; @={(0,0),(0,10),(10,10),(10,0)}; s0="prev" @@{;"prev";**@{-}="prev"}; (0,10); (11.2,19.4)**\dir{.}; (10,0); (11.9,19)**\dir{.}; (2,6)*{\mydot} -(0,1.3)*{\scriptscriptstyle 1}; (8,4)*{\mydot} -(0,1.3)*{\scriptscriptstyle 2}; (6,-1.8)*{\scriptstyle\alpha_{12}=5}; @i; @={(16,0),(16,10),(26,10),(26,0)}; s0="prev" @@{;"prev";**@{-}="prev"}; (16,0); (14.1,15)**\dir{.}; (26,10); (14.7,15.4)**\dir{.}; (18,3)*{\mydot} -(0,1.3)*{\scriptscriptstyle 3}; (24,7)*{\mydot} -(0,1.3)*{\scriptscriptstyle 4}; (22,-1.8)*{\scriptstyle\alpha_{34}=1}; (65,29); (88,29)**\crv{(55,18)&(74,20)&(80,26)&(100,18)&(96,26)}; (68,25)*{X}; (82,26)*{\mydot}; @i; @={(78,12),(78,22),(88,22),(88,12)}; s0="prev" @@{;"prev";**@{-}="prev"}; (78,22); (81.5,26.1)**\dir{.}; (88,22); (82.3,26.2)**\dir{.}; (84.5,15)*{\mydot}; (81.5,19)*{\mydot}; @i; @={(70,0),(70,10),(80,10),(80,0)}; s0="prev" @@{;"prev";**@{-}="prev"}; (70,10); (81.2,19.4)**\dir{.}; (80,0); (81.9,19)**\dir{.}; (72,6)*{\mydot} -(0,1.3)*{\scriptscriptstyle 1}; (78,4)*{\mydot} -(0,1.3)*{\scriptscriptstyle 2}; @i; @={(86,0),(86,10),(96,10),(96,0)}; s0="prev" @@{;"prev";**@{-}="prev"}; (86,0); (84.1,15)**\dir{.}; (96,10); (84.7,15.4)**\dir{.}; (91,5)*{\mydot} @i; @={(86,-12),(86,-2),(96,-2),(96,-12)}; s0="prev" @@{;"prev";**@{-}="prev"}; (86,-2); (90.7,5.2)**\dir{.}; (96,-2); (91.3,5.2)**\dir{.}; (88,-9)*{\mydot} -(0,1.3)*{\scriptscriptstyle 3}; (94,-5)*{\mydot} -(0,1.3)*{\scriptscriptstyle 4}; (110,25)*{\txt{Level 0}}; (110,15)*{\txt{Level 1}}; (110, 5)*{\txt{Level 2}}; (110,-5)*{\txt{Level 3}}; (47,5)*\txt{as $\alpha_{34} \to 0$}; (60,7) \ar @/^0.4ex/ (35,7); \endxy \end{figure} Pictures like the left one in Figure~6 will represent generic points of~$Z$. It~consists of three levels: \begin{enumerate} \renewcommand\theenumi{\arabic{enumi}} \setcounter{enumi}{-1} \item one point in~$X$, \item a screen for $1234$ with two distinct points, and \item a pair of screens $S_{12}$ and $S_{34}$ together with their \newterm{scale factors} $`a_{12}$ and $`a_{34}$, where the pair $[`a_{12} \!:\! `a_{34}]$ is considered as a point in~$\P^1$. \end{enumerate} The scale factors serve to compare the approach speeds of the pairs 12~and~34 by keeping track of independent dilations of their respective screens: for all nonzero scalars~$`f_i$, the pairs $(S_i,`a_i)$ and $(`f_i S_i,`a_i/`f_i)$ are identified, where the screen in the second pair is the dilation of~$S_i$ by the factor of~$`f_i$, as in Figure~5. Nongeneric points of~$Z$, which lie in $Z\cap D^{12}$ and $Z\cap D^{34}$, correspond to incomparable speeds and to the points $[0\!:\!1]$ and $[1\!:\!0]$ in~$\P^1$. They result from collisions mentioned in the beginning of the section. Keeping the screen~$S_{34}$ fixed while letting $`a_{34} \to 0$ is the same thing as keeping $`a_{34}$ fixed while contracting the screen. In the limit the two points in it collide, but a new screen appearing on level~$3$ separates them. Trivial screens, which contain a single point, may be omitted from the pictures. Similarly, points in $D^{12|34}$ away from $D^{1234}$ are represented by configurations of two distinct points in~$X$, labeled 12 and 34, plus screens $S_{12}$ and $S_{34}$ together with their scale factors, generically on the same level and degenerating to two levels. The microscopic levels in Figure~6 correspond to the intersecting divisors: the first to $D^{1234}$, the second to $D^{12|34}$ and, in the right half of the figure, the third to $D^{34}$. Accordingly, trees that link screens together acquire some extra structure: levels of vertices. For example, the two pictures in Figure~6 correspond to the middle and right trees in Figure~1. Such trees index the strata of~$X\pd<4>$. Scale factors are redundant on any level that contains only one nontrivial screen. Since the middle tree in Figure~1 is, up to relabeling, the only tree with four leaves in which two vertices may be on the same level, points in $X\pd<4>$ outside the three additional divisors will have exactly the same screen description as for~$X[4]$. In fact, forgetting the scale factors gives a map~$\vartheta_4 \colon X\pd<4> \to X[4]$ that blows down the divisor $D^{12|34}$ to the stratum $D(12) \cap D(34)$, and respectively for $D^{13|24}$ and $D^{14|23}$. A combinatorial basis is necessary in order to generalize these ideas to an arbitrary number of points, and it is very easy to find. The definition of~$`D^{S}$ for any subset~$S$ of $[n]=\{1,\dots,n\}$ applied to $S=\{k\}$ gives $`D^{\{k\}}=X^n$, hence $`D^{123}=`D^{123} \cap `D^{4}$ and so on. The true combinatorial basis will thus be {\em the partitions of the set\/}~$[n]$. Indeed, when $n=4$, the first blowup is that of $`D=`D^{1234}$, which corresponds to the only partition into one block; the next stage blowup centers correspond to all partitions into two blocks; finally, all those corresponding to partitions into three blocks are blown up: $`D^{12}=`D^{12} \cap `D^{3} \cap `D^{4}$ and so on. \section{Combinatorial background} \label{sec-combinat} This section is a short primer on the language of the rest of the paper: it deals with basic properties of set partitions and a bijection between partition chains and leveled trees. Let $[n]$ denote the set $\{1,\dots,n\}$ of integers. A~partition~$`p$ of~$[n]$ is a set of disjoint subsets of~$[n]$, called the blocks of~$`p$, whose union is~$[n]$. Nonsingleton blocks are called~\newterm{essential}. The two functions of partitions that are most important for this work are $`r(`p)$, the number of blocks, and $`e(`p)$, the number of essential blocks. The integer partition whose parts are {\em one less than} the cardinalities of the essential blocks of~$`p$ is called the \newterm{essential shape} of~$`p$ and denoted by~$`l(`p)$. For example, ${`p_1} = \{12357,9,468\}$ and ${`p_2} = \{15,23,7,9,468\}$ are two partitions of~$[9]$ with \begin{equation*} \begin{align*} {}`r({`p_1}) &= 3, & {}`e({`p_1}) &= 2, & {}`l({`p_1}) &= (4,2), &&\\ {}`r({`p_2}) &= 5, & {}`e({`p_2}) &= 3, & {}`l({`p_2}) &= (2,1,1). && \end{align*} \end{equation*} Let~$\Ln$ be the set of all partitions of $[n]$. There is a refinement partial order on~$\Ln$: ${`p_1}\leqslant{`p_2}$ whenever each block of~${`p_2}$ is contained in a block of~${`p_1}$, as in the example. This makes $\Ln$ a ranked lattice, with $`r(`p)$ being the rank function. The minimal (bottom) and maximal (top) elements of~$\Ln$ are denoted by $\bot$ and $\top$ respectively. The Stirling number of the second kind $S(n,k)$ is the number of partitions of~$[n]$ into exactly~$k$ blocks. Many textbooks on combinatorics discuss these numbers and the partition lattice, for instance, Andrews~\cite{Andrews} and Stanley~\cite{Stanley}. An interval $[`p',`p'']$ in a lattice~$L$ is its subset $\{`p \thinspace|\thinspace `p' \< `p \< `p''\}$. In~$\Ln$, every lower interval $[\bot,`p]$ is isomorphic to $L_{[`r(`p)]}$ and every upper interval $[`p,\top]$ is isomorphic to $L_{[`n_1+1]} \times\dots\times L_{[`n_r+1]}$, where $`l = `l(`p) = (`n_1,\dots,`n_r)$ is the essential shape of~$`p$. This product will be denoted by~$L_`l$. A totally ordered subset of a partially ordered set is called a chain. The length of a chain is the number of its elements. Half of the chains in $\Ln$ contain the top (finest) partition, and the other half do not; from now on, a chain will mean a partition chain of the latter kind. \begin{figure}[b] \label{fig-leveled} \xy (88,-8)*{\mycaption{7in}{ From a partition chain to a leveled tree (and back)}}; (-7,14.7)*{\scriptstyle \pi_1}; (-7, 9.7)*{\scriptstyle \pi_2}; (-7, 4.7)*{\scriptstyle \pi_3}; (16,20)="root" *{\bullet}; (10,15)="a12357" *{\scriptstyle 12357}; (18,15)="a9" *{\scriptstyle 9}; (24,15)="a468" *{\scriptstyle 468}; (4,10)="b15" *{\scriptstyle 15}; (10,10)="b23" *{\scriptstyle 23}; (14.75,10)="b7" *{\scriptstyle 7}; (19.8,10)="b9" *{\scriptstyle 9}; (26.5,10)="b468" *{\scriptstyle 468}; (0.6,5)="c1" *{\scriptstyle 1}; (4.3,5)="c5" *{\scriptstyle 5}; (10,5)="c23" *{\scriptstyle 23}; (15.5,5)="c7" *{\scriptstyle 7}; (20,5)="c9" *{\scriptstyle 9}; (25.5,5)="c46" *{\scriptstyle 46}; (29.8,5)="c8" *{\scriptstyle 8}; (0,0)="d1" *{\scriptstyle 1}; (4,0)="d5" *{\scriptstyle 5}; (8,0)="d2" *{\scriptstyle 2}; (12,0)="d3" *{\scriptstyle 3}; (16,0)="d7" *{\scriptstyle 7}; (20,0)="d9" *{\scriptstyle 9}; (24,0)="d4" *{\scriptstyle 4}; (28,0)="d6" *{\scriptstyle 6}; (32,0)="d8" *{\scriptstyle 8}; "root"; "a12357"+(1,1.2)**\dir{-}; "root"; "a9"+(-0.3,1.2)**\dir{-}; "root"; "a468"+(-1.3,1.2)**\dir{-}; "a12357"-(2,1.2); "b15"+(1,1.2)**\dir{-}; "a12357"-(0,1.2); "b23"+(0,1.2)**\dir{-}; "a12357"-(-2,1.2); "b7"+(-0.5,1.2)**\dir{-}; "a9"-(-0.5,1.2); "b9"+(-0.3,1.2)**\dir{-}; "a468"-(-0.5,1.2); "b468"+(-0.3,1.2)**\dir{-}; "b15"-(1,1.2); "c1"+(0.3,1.2)**\dir{-}; "b15"-(-0.5,1.2); "c5"+(0,1.2)**\dir{-}; "b23"-(0,1.2); "c23"+(0,1.2)**\dir{-}; "b7"-(-0.3,1.2); "c7"+(0,1.2)**\dir{-}; "b9"-(0,1.2); "c9"+(0,1.2)**\dir{-}; "b468"-(0.2,1.2); "c46"+(0.2,1.2)**\dir{-}; "b468"-(-1,1.2); "c8"+(-0.5,1.2)**\dir{-}; "c1"-(0.2,1.2); "d1"+(0,1.2)**\dir{-}; "c5"-(0.1,1.2); "d5"+(0,1.2)**\dir{-}; "c23"-(0.7,1.2); "d2"+(0,1.2)**\dir{-}; "c23"-(-0.7,1.2); "d3"+(0,1.2)**\dir{-}; "c7"-(-0.2,1.2); "d7"+(0,1.2)**\dir{-}; "c9"-(0,1.2); "d9"+(0,1.2)**\dir{-}; "c46"-(0.5,1.2); "d4"+(0,1.2)**\dir{-}; "c46"-(-1,1.2); "d6"+(-0.3,1.2)**\dir{-}; "c8"-(-0.5,1.2); "d8"+(-0.2,1.2)**\dir{-}; (35,13)*{\leftrightarrow}; (39,0)="m"; "m"+(16,20)="root" *{\bullet}; "m"+(10,15)="12357" *{\scriptstyle 12357}; "m"+(26.5,10)="468" *{\scriptstyle 468}; "m"+(4,10)="15" *{\scriptstyle 15}; "m"+(10,5)="23" *{\scriptstyle 23}; "m"+(25.5,5)="46" *{\scriptstyle 46}; "m"+(0,0)="1" *{\scriptstyle 1}; "m"+(4,0)="5" *{\scriptstyle 5}; "m"+(8,0)="2" *{\scriptstyle 2}; "m"+(12,0)="3" *{\scriptstyle 3}; "m"+(16,0)="7" *{\scriptstyle 7}; "m"+(20,0)="9" *{\scriptstyle 9}; "m"+(24,0)="4" *{\scriptstyle 4}; "m"+(28,0)="6" *{\scriptstyle 6}; "m"+(32,0)="8" *{\scriptstyle 8}; "root"; "12357"+(1,1.2)**\dir{-}; "root"; "468"+(-1,1.2)**\dir{-}; "root"; "9"+(0,1.2)**\dir{-}; "12357"-(1.5,1.2); "15"+(1,1.2)**\dir{-}; "12357"-(0,1.2); "23"+(0,1.2)**\dir{-}; "12357"-(-1.5,1.2); "7"+(0,1.2)**\dir{-}; "468"-(0.5,1.2); "46"+(0,1.2)**\dir{-}; "468"-(-0.5,1.2); "8"+(-0.2,1.2)**\dir{-}; "15"-(0.5,1.2); "1"+(0,1.2)**\dir{-}; "15"-(-0.5,1.2); "5"+(0,1.2)**\dir{-}; "23"-(0.7,1.2); "2"+(0,1.2)**\dir{-}; "23"-(-0.7,1.2); "3"+(0,1.2)**\dir{-}; "46"-(0.5,1.2); "4"+(0,1.2)**\dir{-}; "46"-(-1,1.2); "6"+(-0.3,1.2)**\dir{-}; "m"+(35,13)*{\leftrightarrow}; "m"+"root"-(0,0.4)="nroot"; "m"+"12357" ="n12357"; "m"+"468" ="n468"; "m"+"15"+(0,0.4)="n15"; "m"+"23"+(0,0.8)="n23"; "m"+"46"+(0,0.8)="n46"; "m"+"1"+(0,1.2) ="n1"; "m"+"2"+(0,1.2) ="n2"; "m"+"3"+(0,1.2) ="n3"; "m"+"4"+(0,1.2) ="n4"; "m"+"5"+(0,1.2) ="n5"; "m"+"6"+(0,1.2) ="n6"; "m"+"7"+(0,1.2) ="n7"; "m"+"8"+(0,1.2) ="n8"; "m"+"9"+(0,1.2) ="n9"; @={"nroot","n12357","n15","n1","n15","n5","n15","n12357","n23","n2", "n23","n3","n23","n12357","n7","n12357","nroot","n9","nroot", "n468","n46","n4","n46","n6","n46","n468","n8","n468","nroot"}; s0="prev" @@{;"prev";**@{-}="prev"}; "nroot" *{\bullet}; "n12357"*{\scriptscriptstyle\bullet}; "n468"*{\scriptscriptstyle\bullet}; "n15"*{\scriptscriptstyle\bullet}; "n23"*{\scriptscriptstyle\bullet}; "n46"*{\scriptscriptstyle\bullet}; "n1"-(0,1.2) *{\scriptstyle 1}; "n2"-(0,1.2) *{\scriptstyle 2}; "n3"-(0,1.2) *{\scriptstyle 3}; "n4"-(0,1.2) *{\scriptstyle 4}; "n5"-(0,1.2) *{\scriptstyle 5}; "n6"-(0,1.2) *{\scriptstyle 6}; "n7"-(0,1.2) *{\scriptstyle 7}; "n8"-(0,1.2) *{\scriptstyle 8}; "n9"-(0,1.2) *{\scriptstyle 9}; (114,20)*\txt{\tiny level 0}; (114,15.33)*\txt{\tiny level 1}; (114,10.66)*\txt{\tiny level 2}; (114, 6)*\txt{\tiny level 3}; \endxy \end{figure} Lengyel represented~\cite{Lengyel} partition chains as trees. If $`g = \{`p_1,\dots,`p_k\}$, where $`p_i < `p_{i+1}$ for $1 \< i \< k$, then the associated tree has the blocks of each partition as its interior vertices, one additional vertex (the root) and leaves labeled by $1,\dots,n$. Edges indicate inclusions of blocks of $`p_{i+1}$ into those of~$`p_i$ and of the elements of~$[n]$ into the blocks of~$`p_k$; they also connect the blocks of~${`p_1}$ to the root. The left tree in Figure~7 goes with the chain $`g = \{`p_1,`p_2,`p_3\}$, where $$ `p_1 = \{12357,9,468\}, \quad `p_2 = \{15,23,7,9,468\}, \quad `p_3 = \{1,5,23,7,9,46,8\}. $$ The $2$-valent vertices (except for the root if it happens to be such) may be called the \newterm{phantom vertices} because it is often convenient to omit them; this gives trees like the middle one in the same figure. Furthermore, labels of interior vertices are also unnecessary. In thus simplified tree the set of interior vertices is the set $\{12357,468,23,46,15\}$ of all essential blocks in the three partitions, and they appear to be on different {\em levels} reflecting how far in the chain they survive unsubdivided. This leads to the following \begin{defn} A \newterm{$k$-leveled tree} is a pair $(T,`y)$, where~$T$ is a rooted tree without $2$-valent vertices, except possibly for the root, and~$`y$ is a surjective poset map from the set of vertices of~$T$ with the parent-descendant partial order to the set of integers $\{0,\dots,k\}$ with its standard order. (The root goes to~$0$.) The number~$`y(v)$ is called the \newterm{level} of the vertex~$v$. The map from leveled trees with marked leaves to usual rooted trees with marked leaves by $(T,`y)\mapsto T$ is denoted by~$`h$. \end{defn} The term {\em leveled tree} belongs to Loday, although his trees are binary~\cite{Loday}. An~inspiring picture evinces that Tonks used leveled trees implicitly~\cite{Tonks}. In~both references the leaves are not marked. The sole purpose of the root is to simplify wording: without it, we would be dealing not only with trees, but also with groves (disjoint unions of trees). The example above demonstrates how to pass from a $k$-chain~$`g$ of partitions of~$[n]$ to a $k$-leveled tree $(T_`g,`y_`g)$ with $n$~marked leaves; this is actually a bijection when restricted to such chains~$`g$ that $\top \not\in `g$. There is a unique (shortest) path from the root of $(T,`y)$ to each leaf, and each pair of such separate at a vertex on certain level~$j$. The labels of the two leaves will be in the same block in the partitions~$`p_i$ for~$i\<j$, and they will be in different blocks in~$`p_i$ for~$i>j$. This defines the $k$-chain $`g(T,`y)$. It will also be useful to associate with a $k$-leveled tree $(T,`y)$ a sequence $\{`l_i(T,`y)\} = \{`l_i(`g)\}$ of integer partitions as follows. While $`l_0$ has just one part, equal to the valency of the root of~$T$, the partition~$`l_i$, $1 \< i \< k$, is to have as many parts as there are vertices of $(T,`y)$ on level~$i$, and each part is to be one less than the number of direct descendants of the corresponding vertex. With that, $`r(`p_1) = `l_0(`g)$ and $[`p_i,`p_{i+1}] \simeq L_{`l_i(`g)}$ for $1 \< i \< k$, where $`g = `g(T,`y) = \{`p_1,\dots,`p_k\}$ and $`p_{k+1} = \top$. For the example above, $`l_0 = (3)$, $`l_1 = (2)$, $`l_2 = (1,1)$ and $`l_3 = (1,1)$. \section{Polyscreens and colored screens} \label{sec-screens} \def\mathbf S{\mathbf S} Partition chains and leveled trees of the previous section play in the polydiagonal compactification $\Xn$ the same role as nests of subsets of~$[n]$ and usual trees (groves) do in the Fulton--MacPherson compactification $X[n]$. They index the strata and are an integral part of the geometric description of points in $\Xn$, explained in this section without any proofs. It is implied by the technical work of Section~\ref{sec-strata}. For a chain $`g = \{`p_1,\dots,`p_k\}$, each point~$\x$ in the stratum~$S_`g$ of~$\Xn$ is represented by a configuration~$\x'$ of distinct points in~$X$ labeled by the blocks of~${`p_1}$ and a coherent sequence of polyscreens $\PS^{`p_1},\dots,\PS^{`p_k}$ at~$\x'$. Let $p(\x',`b)$ be the point in the configuration~$\x'$ labeled by the block of~${`p_1}$ that contains $`b \subseteq [n]$. This makes sense for every block~$`b$ of every $`p \> `p_1$. \begin{defn} A \newterm{polyscreen} $\PS^{`p}$ at~$\x'$ is given by: for each block $`b_i$ of~$`p$, a configuration~$\mathbf S_i$ of $\mbox{card}(`b_i)$ points in the tangent space to~$X$ at $p(\x',`b_i)$, labeled by the elements of~$`b_i$, and a nonzero scalar~$`a_i$, called the \newterm{scale factor} of~$\mathbf S_i$. The data is considered modulo the following relations: \begin{enumerate} \item translation of any screen~$\mathbf S_i$; \item dilation of any screen~$\mathbf S_i$ with compensating change of its scale factor: $(\mathbf S_i,`a_i)\sim(`f\mathbf S_i,`f^{-1}`a_i)$, $`f\in\kt$; \item simultaneous multiplication of all~$`a_i$ by an element of~$\kt$ (rescaling). \end{enumerate} A sequence of polyscreens $\PS^{`p_1},\dots,\PS^{`p_k}$ is \newterm{coherent} if, for all $j = 1,\dots,k-1$, two labeled points in $\PS^{`p_j}$ coincide if and only if their labels belong to the same block of $`p_{j+1}$, and all labeled points in $\PS^{`p_k}$ are distinct. \end{defn} Coherence makes a sequence $\PS^{`g}$ conform to the leveled tree $(T_`g,`y_`g)$, as the example in Figure~8 of a point in $X\pd<9>$ does to the (right) tree in Figure~7. This means that the root of the tree corresponds to~$X$, each internal vertex has a screen attached to it and the direct descendants of each vertex form a configuration of distinct points in~$X$ or in the respective screen. The screens in $\PS^{`g}$ attached to the phantom vertices of $(T_`g,`y_`g)$ contain just one distinct labeled point and are called \newterm{trivial}; they carry no information and are left out of the pictures. \begin{figure} \label{fig-polyscreens} \xy (-40,0)*{\ }; (-20,0)*{\mycaptionempty}; (-5,29); (48,29)**\crv{(-15,18)&(4,20)&(10,26)&(55,18)&(46,26)}; (-2,25)*{X}; (12,26)="12357" *{\mydot}; (23,28) *{\mydot} +(1.3,0)*{\scriptscriptstyle 9}; (40,24)="468" *{\mydot}; (10,0)="screenwd", (0,10)="screenht", (5,5)="gocenter", (7.4,-1.6)="golabel", (8,12)="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "lu"; "12357"+(-0.5,0.1)**\dir{.}; "ru"; "12357"+(0.3,0.2)**\dir{.}; "center"+(2,3.5)*{\mydot} +(1.3,0)*{\scriptscriptstyle 7}; "center"-(2.5,1)="15"*{\mydot}; "center"-(0,3)="23"*{\mydot}; (0,0)="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "lu"; "15"+(-0.3,0.4)**\dir{.}; "rd"; "15"+(0.4,0)**\dir{.}; "center"-(3,-1)*{\mydot} -(0,1.3)*{\scriptscriptstyle 1}; "center"+(3,-1)*{\mydot} -(0,1.3)*{\scriptscriptstyle 5}; "ld"+"golabel"*{\scriptstyle\alpha=1}; (14,-12)="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "ld"; "23"+(-0.4,0)**\dir{.}; "ru"; "23"+(0.2,0.4)**\dir{.}; "center"-(3,2)*{\mydot} -(0,1.3)*{\scriptscriptstyle 2}; "center"+(3,2)*{\mydot} -(0,1.3)*{\scriptscriptstyle 3}; "ld"+"golabel"*{\scriptstyle\alpha=4}; (36,0)="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "lu"; "468"+(-0.5,0.1)**\dir{.}; "ru"; "468"+(0.3,0.2)**\dir{.}; "center"+(3,3)*{\mydot} +(1,-0.6)*{\scriptscriptstyle 8}; "center"-(2,2)="46"*{\mydot}; "ld"+"golabel"*{\scriptstyle\alpha=2}; (28,-12)="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "lu"; "46"+(-0.5,0.1)**\dir{.}; "rd"; "46"+(0.3,0.2)**\dir{.}; "center"-(3,1)*{\mydot} +(0.9,-1)*{\scriptscriptstyle 6}; "center"+(3,1)*{\mydot} +(0.8,-0.8)*{\scriptscriptstyle 4}; "ld"+"golabel"*{\scriptstyle\alpha=1}; (65,27)*{\txt{Level 0}}; (65,16)*{\txt{Level 1}}; (65, 5)*{\txt{Level 2}}; (65,-6)*{\txt{Level 3}}; \endxy \end{figure} Nontrivial screens in a polyscreen $\PS^{`p_j}$ are exactly Fulton--MacPherson screens for those blocks of~$`p_j$ that are subdivided in $`p_{j+1}$ (all essential blocks of~$`p_j$ if $j=k$). The data of $\PS^{`p_j}$ is equivalent to this collection of screens together with the point in the projective space $\P^{r_j-1}$ given by the $r_j$-tuple of scale factors, where $r_j$ is the number of nontrivial screens in~$\PS^{`p_j}$. If $`g = \{`p_1,\dots,`p_k\}$ starts with the bottom partition $\bot$, then for all points~$\x$ in~$S_`g$ the configuration~$\x'$ is a single point~$p$ in~$X$, and all screens in $\PS^`g(\x)$ are based on the {\em same} tangent space $T_p X$. Under the additional assumption that $\operatorname{char}\Bbbk = 0$, now made for the rest of this section, the data of each polyscreen $\PS^{`p_j}(\x)$ then fits into a single {\em colored screen} $\CS^{`p_j}(\x)$. \begin{defn} Let a \newterm{color} be any nonempty subset of~$[n]$. A \newterm{colored screen} $\CS^`p$ at~$p$ is a configuration of $n$~colored points $x_1,\dots,x_n$ in~$T_p X$, considered modulo dilations of~$T_p X$, where the color of~$x_i$ is the block of~$`p$ that contains~$i$, such that the points of each color are centered around the origin (their vector sum is~$0$). A sequence $\CS^{`p_1},\dots,\CS^{`p_k}$ is \newterm{coherent} if, for all $j = 1,\dots,k-1$, two points of the same color coincide in $\CS^{`p_j}$ if and only if they have the same color in $\CS^{`p_{j+1}}$, and in $\CS^{`p_k}$ no points of the same color coincide. \end{defn} To convert a polyscreen $\PS^`p$ into a colored screen $\CS^`p$, first translate the representative screens of $\PS^`p$ to center the points around the origin, then dilate them to make all scale factors equal. Identifying now the underlying spaces of the screens, place several configurations in the same~$T_p X$. To tell them apart, colors of points are added as a way of recording which one of the screens each point comes from. Figure~9 shows the simplest nontrivial example. \begin{figure}[t] \label{fig-conversion} \xy (-13,0)*{\ }; (75,-8)*{\mycaption{4in}{Conversion to color}}; (14,0)="screenwd", (0,14)="screenht", (7,7)="gocenter", (10,-1.8)="golabel", (24,0)="stepright", (0,0)="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "center"-(2,-2)*{\mydot} -(0,1.3)*{\scriptscriptstyle 1}; "center"+(2,-2)*{\mydot} -(0,1.3)*{\scriptscriptstyle 2}; "ld"+"golabel"*{\scriptstyle\alpha_{12}=2}; (3,0)+"rd"="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "center"-(3,1.5)*{\mydot} -(0,1.3)*{\scriptscriptstyle 3}; "center"+(3,1.5)*{\mydot} -(0,1.3)*{\scriptscriptstyle 4}; "ld"+"golabel"-(1,0)*{\scriptstyle\alpha_{34}=-1}; "rd"+(5,6)*{\mapsto}; "stepright"+"ld"="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "center"-(4,-4)*{\mydot} -(0,1.3)*{\scriptscriptstyle 1}; "center"+(4,-4)*{\mydot} -(0,1.3)*{\scriptscriptstyle 2}; "ld"+"golabel"*{\scriptstyle\alpha_{12}=1}; (3,0)+"rd"="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "center"+(3,1.5)*{\mydot} -(0,1.3)*{\scriptscriptstyle 3}; "center"-(3,1.5)*{\mydot} -(0,1.3)*{\scriptscriptstyle 4}; "ld"+"golabel"*{\scriptstyle\alpha_{34}=1}; "rd"+(5,6)*{\mapsto}; "stepright"+"ld"="ld", "ld"+"screenht"="lu", "ld"+"screenwd"="rd", "ld"+"screenht"+"screenwd"="ru", "ld"+"gocenter"="center", @i; @={"ld","lu","ru","rd"}; s0="prev" @@{;"prev";**@{-}="prev"}; "center"-(4,-4)*{\mydot} -(0,1.3)*{\scriptscriptstyle 1}; "center"+(4,-4)*{\mydot} -(0,1.3)*{\scriptscriptstyle 2}; "center"+(3,1.5)*{\scriptstyle\circ} -(0,1.3)*{\scriptscriptstyle 3}; "center"-(3,1.5)*{\scriptstyle\circ} -(0,1.3)*{\scriptscriptstyle 4}; \endxy \end{figure} Since this conversion of polyscreens into colored screens respects coherence, points in~$\Xn$ corresponding to collisions at a single point in~$X$ can be viewed in terms of coherent sequences of colored screens. This interpretation is useful in Section~\ref{sec-isotropy} for studying the natural action of~$\Sym_n$ on~$\Xn$. \section{Construction of the compactification} \label{sec-construct} For a partition~$`p$ of~$[n]$, denote by $`D^`p\subseteq X^n$ the subset of all points $(x_1,\dots,x_n)$ with $x_i=x_j$ whenever $i$ and~$j$ are in the same block of~$`p$, and call $`D^`p$ a \newterm{ polydiagonal}. The diagonals of~$X^n$ correspond to partitions with only one essential block. The set of all polydiagonals in~$X^n$ is naturally a lattice isomorphic to $\Ln$, with its top element~$X^n$ itself. \begin{thm} \label{construction} The following\/ $(n-1)$-stage sequence of blowups results in a smooth compactification $\Xn$ of the configuration space of\/~$n$ distinct labeled points in a~smooth algebraic variety~$X$: \begin{itemize} \item the first stage is the blowup of\/~$`D$, the small diagonal of\/~$X^n$; \item the $k$-th stage, $1<k<n$, is the blowup of the disjoint union of the previous stage proper transforms $Y^`p_{k-1}$ of\/ $`D^`p$, for all partitions~$`p$ of the set\/ $[n]=\{1,\dots,n\}$ into exactly\/~$k$ blocks. \end{itemize} \end{thm} \begin{rem} In the language of De~Concini and Procesi~\cite{DeCP}, the building set for this iterated blowup construction consists of {\em all} possible intersections of the diagonals of~$X^n$, and therefore it is maximal. The building set of the Fulton--MacPherson compactification includes only those intersections that fail to be normal, so $X[n]$ is the minimal compactification of $\Conf(X,n)$ with the property that the complement to the configuration space is a divisor with normal crossings. \end{rem} Two smooth subvarieties $U$~and~$V$ of a smooth algebraic variety~$W$ are said to \newterm{intersect cleanly} if $U\not\subset V\not\subset U$, their scheme-theoretic intersection is smooth and the tangent bundles satisfy $T(U\cap V)=TU\cap TV$. Two polydiagonals $`D^{`p_1}$ and $`D^{`p_2}$ in~$X^n$ intersect cleanly unless one of them contains the other; the noncontainment condition is that the partitions ${`p_1}$~and~${`p_2}$ are incomparable in~$\Ln$. Recall two standard results about the behaviour of clean intersections under blowups: \begin{lem} \label{clean} Let\/~$W$ be a smooth algebraic variety and let\/ $U$,~$V$ be smooth subvarieties of\/~$W$ intersecting cleanly. Then \begin{enumerate} \item \label{be-disjoint} the proper transforms of\/ $U$ and\/~$V$ in $\Bl_{U\cap V}W$ are disjoint; \item \label{ZinUcapV} if\/~$Z$ is a smooth subvariety of~$U\cap V$, then the proper transforms of\/ $U$ and\/~$V$ in $\Bl_Z W$ intersect cleanly.\qed \end{enumerate} \end{lem} \begin{proof}[Proof of Theorem 1.] Denote the space obtained at stage~$k$ by~$Y_k$ and organize the projections of the fiber squares of all stages as \begin{equation} \label{sequence} \xymatrix @-3pt { \Xn= Y_{n-1}\ar[r] & Y_{n-2}\ar[r] & {}\dots\ar[r] & Y_1 \ar[r] & Y_0=X^n. } \end{equation} Then $Y^`p_0=`D^`p$ and $Y^`p_k$ is the proper transform of~$Y^`p_{k-1}$ in~$Y_k^{\vphantom{`p}}$ if $`r(`p)\ne k$, while $Y^`p_{`r(`p)}$ is the component of the exceptional divisor over $Y^`p_{`r(`p)-1}$. The statement will follow once it has been shown that the stated sequence of blowups can indeed be performed. For this, it suffices to check that the centers of those simultaneous blowups will have indeed become disjoint after the previous stages of the construction. The proof will be done by induction on~$k$, for all $\Xn$ at the same time; after stage~$k$, the induction will stop for $X\pd<k+1>$, and it will continue on for $\Xn$ with $n>k+1$. For any pair of distinct partitions ${`p_1}$ and~${`p_2}$ of~$[n]$ into two blocks, their meet ${`p_1}\wedge{`p_2}$ is the \inquotes{nonpartition}\negthickspace, so $`D^{`p_1}\cap`D^{`p_2}=`D^{{`p_1}\wedge{`p_2}}=`D$, the small diagonal of~$X^n$. By Lemma~\ref{clean}\ref{be-disjoint}, the transforms $Y_1^{`p_1}$ and $Y_1^{`p_2}$ will be disjoint, making the second stage possible. Assume that stage $k-1$ has been performed; this means that the varieties $\Xn$ have been constructed for $1\leqslant n\leqslant k$, and only those for $n>k$ are still being built. Also assume that the proper transforms $Y^`p_{k-1}$ for~$`p$ with $`r(`p)=k$ are disjoint. For each partition $`p\in \Ln$ with $`r(`p)=k$, the projection $X\pd<k>\to X^k$ pulls back the obvious isomorphism $X^k \simeq `D^`p \subset X^n$ to an isomorphism $X\pd<k> \simeq Y_{k-1}^`p \subset Y_{k-1}^{\vphantom{`p}}$. All these subvarieties are disjoint by the inductive assumption, and can all be blown up at the same time. This defines the variety $X\pd<k+1>$. To provide the inductive step necessary to continue the construction of $\Xn$ for $n>k+1$, the intersection $Y_k^{`p_1}\cap Y_k^{`p_2}$ must be empty for all pairs of distinct ${`p_1}$, ${`p_2}$ in $\Ln$ with $`r({`p_1})=`r({`p_2})=k+1$. Such a pair automatically satisfies the noncontainment condition; so $`D^{`p_1}\cap`D^{`p_2}=`D^{{`p_1}\wedge{`p_2}}$ is a clean intersection. Since $`r=`r({`p_1}\wedge{`p_2})<k+1$, a repeated use of Lemma~\ref{clean}\ref{ZinUcapV} shows that $Y^{`p_1}_{`r-1}\cap Y^{`p_2}_{`r-1}=Y^{{`p_1}\wedge{`p_2}}_{`r-1}$ is a clean intersection, and then Lemma~\ref{clean}\ref{be-disjoint} implies that $Y^{`p_1}_`r\cap Y^{`p_2}_`r$ is empty. The proper transforms of $`D^{`p_1}$ and $`D^{`p_2}$ become disjoint after stage $`r\<k$, and the proof is complete. \end{proof} \begin{corol} \label{middleX} For each $`p\in\Ln$ we have $Y_{`r(`p)-1}^`p\simeq X\pd<`r(`p)>$. \end{corol} \begin{proof} This has been obtained while proving the theorem, and is formulated separately only for the ease of future reference. \end{proof} \begin{FBL} Let\/ $V^1_0\subset V^2_0\subset\dots\subset V^s_0\subset W_0^{\vphantom{s}}$ be a flag of smooth subvarieties in a smooth algebraic variety~$W_0$. For $k=1,\dots,s$, define inductively: $W_k$ as the blowup of\/ $W_{k-1}$ along $V^k_{k-1}$; $V^k_k$ as the exceptional divisor in~$W_k$; and\/ $V^i_k$, for $i\ne k$, as the proper transform of\/~$V^i_{k-1}$ in~$W_k$. Then the preimage of\/~$V^s_0$ in the resulting variety~$W_s$ is a normal crossing divisor $V_s^1\cup\dots\cup V_s^s$. \end{FBL} \begin{rem} This auxiliary result is implicit in earlier works \cite{FM,Kapranov:Chow}. \end{rem} \begin{proof} In a blowup $p \colon \Bl_Z W \to W$ of a smooth algebraic variety~$W$ along a smooth center~$Z$, if~$\tilde{V}$ is the proper transform of a smooth variety~$V \supset Z$, then in terms of ideal sheaves $\I(p^{-1}(V))=\I(\tilde{V})\cdot\I(E)$. Applied at each step, this equality yields $\I(p_s^{-1}(V^s_0)) = \I(V^1_s) \times\dots\times \I(V^s_s)$, where $p_s\colon W_s\to W_0$ denotes the composition of the stated blowups. \end{proof} \begin{prop} \label{divisor} For each partition~$`p$ of\/~$[n]$ with at least one essential block, there is a smooth divisor $D^`p\subset \Xn$ such that: \begin{enumerate} \item The union of these divisors is $D=\Xn\smallsetminus \Conf(X,n)$. \item Any set of these divisors meets transversally. \item An intersection $D^{`p_1}\cap\dots\cap D^{`p_k}$ of divisors is nonempty exactly when the partitions form a chain. In other words, the incidence graph of\/~$D$ coincides with the comparability graph of the lattice~$\Ln$ with the top partition removed. \end{enumerate} \end{prop} \begin{corol} \label{strata} \begin{enumerate} \item $\Xn$ is stratified by strata $\,S_`g=\bigcap_{`p\in`g}D^`p$ parametrized by all chains~$`g$ in~$\Ln$. \item The codimension of\/~$S_`g$ in~$\Xn$ is equal to the length of\/~$`g$. \item The intersection of two strata $S_`g$~and\/~$S_{`g'}$ is nonempty exactly when $`g\cup`g'$ is a chain, in which case $S_`g\cap S_{`g'}=S_{`g\cup`g'}$. In particular, $S_`g\supset S_{`g'}$ if and only if\/~$`g\subset`g'$. \end{enumerate} \end{corol} \begin{proof} We concentrate on the normal crossing property, which implies the other claims. By construction, the proper transform of every polydiagonal $`D^`p\subset X^n$ under $\Xn\to X^n$ is a smooth divisor; it will be denoted by~$D^`p$. The proper transforms of $`D^{`p_1}$ and $`D^{`p_2}$ become disjoint when that of their intersection $`D^{{`p_1}\wedge{`p_2}}$ is blown up, unless one of $`D^{`p_1}$ and $`D^{`p_2}$ contains the other, that is, unless $\{{`p_1},{`p_2}\}$ is a chain. In order to show that for any saturated (maximal length) chain $`g=\{`p_i\}$, the union $D^`g=D^{`p_1}\cup\dots\cup D^{`p_{n-1}}$ is a normal crossing divisor in~$\Xn$, consider the flag of polydiagonals $`D^{{`p_1}}\subset\dots\subset`D^{`p_{n-1}}\subset X^n$. The blowups of $Y^`p_{`r(`p)-1}$ for $`p\not\in`g$ are irrelevant for the intersection of the components of~$D^`g$ because their centers are disjoint from $$ \bigcap_{i=1}^{n-1} Y^{`p_i}_{`r(`p)-1}; $$ hence, the Flag Blowup Lemma can be applied. The normal crossing property of $D^`g$ follows by the lemma, and so does the proposition: since any chain~$`g'$ is refined by a saturated chain~$`g$, the components of $\bigcup_{`p\in`g'} D^`p$ form a subset of components of $\bigcup_{`p\in`g} D^`p$. \end{proof} \subsection*{Enumeration of the strata.} The number of strata in~$\Xn$, $n>1$, is equal to the number $2Z(n)$ of chains in~$\Ln$. There is a factor of $2$ here because half of the chains contain $\bot$ and half do not (the top~$\top$ is always excluded). Sloane and Plouffe~\cite{EIS} catalogued the sequence $\{Z(n)\}$ of integers as M3649. Since the following recurrence relation is immediate: $$ \label{recurrenceZn} Z(n)=\sum_{k=1}^{n-1}S(n,k)Z(k), $$ the first few values of $Z(n)$ are easy to compute. No closed general formula is known, although Babai and Lengyel described the asymptotics of $Z(n)$, up to yet undetermined constant \cite{BabaiLengyel,Lengyel}. Here is a small table of the numbers of strata in $X[n]$ and $\Xn$: $$ \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline $n$ &2 &3 &4 &5 &6 &7 &8 &9 \\ \hline $X[n]$ &2 &8 &52 &472 &5504 &78416 &1320064 &25637824 \\ \hline $\Xn$ &2 &8 &64 &872 &18024 &525520 &20541392 &1036555120 \\ \hline \end{tabular} $$ As codimension-$1$ strata are the components~$D^`p$ of the divisor at infinity, there are $B(n)-1$ of them, where $B(n)$ is the Bell number, equal to the number of partitions of~$[n]$. The minimal strata have codimension $n-1$ and correspond to saturated chains in~$\Ln$, whose number is $2^{1-n}n!(n-1)!$. \section{The Hodge polynomial\ of $\Xn$} \label{sec-hodge} If~$X$ is a smooth complex algebraic variety, the construction of~$\Xn$ allows an easy derivation of a formula for the Hodge polynomial, hence, for the Poincar\'e polynomial\ of~$\Xn$ in terms of those of~$X$. The notion of a {\em virtual Poincar\'e polynomial\/} extends the usual one to all complex algebraic varieties and provides a good tool for computing the Poincar\'e polynomial s of blowup constructions. \begin{lem} \label{popo} \begin{enumerate} \item If\/~$Y$ is smooth and compact, the virtual Poincar\'e polynomial\ $P(Y)$ coincides with the usual Poincar\'e polynomial\ of\/~$Y$. \item If\/ $Z$ is a closed subvariety of\/ $Y$, then $P(Y)=P(Z)+P(Y\smallsetminus Z)$. \item If\/ $Y'\to Y$ is a bundle with fiber $F$ which is locally trivial in the Zariski topology, then $P(Y')=P(Y)P(F)$. \qed \end{enumerate} \end{lem} Using Deligne's mixed Hodge theory \cite{Deligne:Hodge,Deligne:poids}, Danilov and Khovanskii defined a refinement of~$P(X)$, the {\em virtual Hodge polynomial\/} $e(X)$, also called the {\em Serre polynomial}, and proved~\cite{DanilovKhovanskii} that it has the properties listed in Lemma~\ref{popo}, also independently found by Durfee~\cite{Durfee}. Cheah, Getzler and Manin computed the Hodge polynomial s of the Fulton--MacPherson compactifications via generating functions \cite{Cheah,Getzler,Manin}, while the original paper dealt with summation over trees (groves). \begin{prop} For any two positive integers $m$~and\/~$n$, there is a polynomial\/ $U^m_n(t,x)$ such that for any smooth $m$-dimensional complex algebraic variety~$X$ the Hodge polynomial\ of $\Xn$ is $e(\Xn;z,\bar{z})=U^m_n(z\bar{z},e(X;z,\bar{z}))$, and in particular, $P(\Xn;t)=U^m_n(t,P(X;t))$. The polynomials $U^m_n(t,x)$ satisfy the recurrence relation $$ U^m_n(t,x) = x^n + \sum_{k=1}^{n-1} S(n,k) h_{(n-k)m}(t) U^m_k(t,x), $$ where $h_d(t)=P(\C\P^{d-1})-1=t^{2d-2}+\dots+t^4+t^2$. \end{prop} \begin{proof} Straightforwardly from the construction of~$\Xn$ and Lemma~\ref{popo}, $$ e(Y_k) = e(Y_{k-1}) + \sum_{`r(`p)=k} \big(e(\P(N^`p))-1\big) e(Y^`p_{k-1}), $$ where~$N^`p$ is the fiber of the normal bundle to $Y^`p_{k-1}$ in~$Y^{\phantom{`p}}_{k-1}$, which is $\C^{(n-k)m}$ by an easy dimension count. Corollary~\ref{middleX} converts this formula into $$ e(Y_k) = e(Y_{k-1}) + S(n,k) h_{(n-k)m}(z\bar{z}) e(X\pd<k>). $$ Since $Y_0=X^n$ and $Y_{n-1}=\Xn$, there results a recurrence relation $$ e(\Xn)=e(X)^n + \sum_{k=1}^{n-1} S(n,k) h_{(n-k)m}(z\bar{z}) e(X\pd<k>), $$ and both claims immediately follow. \end{proof} A nonrecursive expression for $U^m_n(t,x)$ can be found by expanding in the right-hand side of the recurrence the terms with the highest~$k$ present in a loop down to $k=2$ terms: $$ U^m_n(t,x)= x^n+ \sum_{s=1}^{n-1} \Bigg(\! x^s \sum_{r=1}^{n-s} \sum_{\mathrm J_{s,n}^{r}} \prod_{i=1}^{r} S(j_i,j_{i-1}) h_{(j_{i}-j_{i-1})m}(t) \!\Bigg), $$ where $\mathrm J_{s,n}^{r}=\{(j_0,\dots,j_r)\in`Z^{r+1} \thinspace|\thinspace s=j_0<\dots< j_r=n\}$. Similar computations of the Hodge polynomial s of the strata in the stratification of~$\Xn$ from Corollary~\ref{strata} can be carried out using the description of their structure given in Section~\ref{sec-strata}. \section{$\Xn$ as a closure and a surjection $\Xn\to X[n]$} \label{sec-closure} In this section I present~$\Xn$ as the closure of the configuration space embedded in a product of blowups, exhibit a surjection $\Xn\to X[n]$, and write it as an iterated blowup. First, the results of Section~\ref{sec-construct} about the structure of~$\Xn$ at infinity should be rephrased in terms of ideal sheaves. Let $\I(`D^`p)$ be the ideal sheaf of $`D^`p$ in~$\O_{X^n}$. For any~$k$, $1\leqslant k\leqslant n-1$, let $`t_k\colon Y_k\to X^n$ be the appropriate composition of projections from Eq.~(\ref{sequence}), let $\I_k(`p)$ be the ideal sheaf in $\O_{Y_k}$ generated by $`t^*_k(\I(`D^`p))$, and also let $\I(Y^`p_k)$ be the ideal sheaf of $Y^`p_k$ in~$\O_{Y_k}$. This notation, although similar to Fulton and MacPherson's, is not quite the same. The assertions of Proposition~\ref{divisor} can be restated as $$ \I_{n-1}(`p)=\prod_{`p'\leqslant`p}\I(D^{`p'}), $$ while at the intermediate stages $$ \I_{k}(`p)=\prod_{`p'\leqslant`p\text{ with }`r(`p')\leqslant k} \I(Y^{`p'}_k). $$ Since $Y^{`p'}_k\subset Y^{\vphantom{`p'}}_k$ is a divisor if $`r(`p')<k$, it follows that $\I_{k}(`p)=\I(Y^{`p}_k)\cdot\mathcal J$, where~$\mathcal J$ is an invertible ideal sheaf. \begin{prop} \label{closure} The variety~$\Xn$ constructed by blowing up is the closure of the configuration space $\Conf(X,n)$ in\/ $$ \prod_{`p\in\Ln}\Bl_{`D^`p}X^n. $$ \end{prop} \begin{rem} The top partition contributes the factor $X^n$. \end{rem} \begin{proof} By induction on~$k$, each $Y_k$ is the closure of~$\Conf(X,n)$ in $$ X^n\times \prod_{`r(`p)\<k}\Bl_{`D^`p}X^n. $$ The basis is clear: $Y_0=X^n$. Then,~$Y_k$ is the blowup of $Y_{k-1}$ along $$ \coprod_{`r(`p)=k} Y^`p_{k-1}, $$ or in other terms, along $$ \I\Bigg(\coprod_{`r(`p)=k}Y^`p_{k-1}\Bigg)=\prod_{`r(`p)=k}\I(Y^`p_{k-1}). $$ This ideal sheaf becomes $$ \I_{k-1}=\prod_{`r(`p)=k}\I_{k-1}(`p) $$ upon multiplying by an invertible ideal sheaf, and blowing up $\I_{k-1}$ is equivalent to taking the closure of the graph of the rational map from $Y_{k-1}$ to $$ \prod_{`r(`p)=k}\Bl_{`D^`p}X^n. $$ This provides the inductive step, and eventually $Y_{n-1}=\Xn$. \end{proof} Both the statement and its proof parallel those by Fulton and MacPherson\ \cite[Prop.~4.1]{FM}, who use pullbacks by $X[n]\to X^n\to X^S$, for $S\subset[n]$, $\#S>1$, and also by $f_S\colon Y_k\to X^n\to X^S$ at the intermediate stages, while here $`t_k\colon Y_k\to X^n$. A slight reformulation of their characterization of~$X[n]$ as a closure elucidates its similarity with $\Xn$. For each~$S$ as before, take the diagonal $`D^S\subset X^n$ and pull back its ideal sheaf by the first of the two arrows whose composition is~$f_S$; this gives the same ideal sheaf $f^*_S(\I(`D))$. \begin{prop} \label{reformulated} The variety $X[n]$ is the closure of $\Conf(X,n)$ in $$ X^n\times\prod_{S\subset[n],\#S>1}\Bl_{`D^S}X^n. $$ \end{prop} The two compactifications can now be related. \begin{prop} \label{ontoFM} For each~$n \> 1$, there is a surjection $\vartheta_n\colon \Xn\to X[n]$. \end{prop} \begin{proof} Start with notation for the products from Propositions \ref{closure}~and~\ref{reformulated}: $$ \Pi = \prod_{`e(`p)\geqslant 1} \Bl_{`D^`p}X^n, \quad \text{ and }\quad \Pi' = \prod_{`e(`p)=1} \Bl_{`D^`p}X^n, $$ where $`e(`p)$ is the number of essential blocks in a partition~$`p$. If~$S$ is the only essential block of~$`p$, then $`D^S=`D^`p$, so $\Pi'$ can indeed be used for $X[n]$. \begin{equation*} \mbox{ \xymatrix @-4mm{ & & {{}\Pi} \ar^p[dd] \\ {}\Conf(X,n) \ar[r] \ar@/^/[urr] \ar@/_/[drr] & X^n \ar@{.>}_{\!\!\!\!\phi}[ur] \ar@{.>}^{\!\!\!\!\psi}[dr] & \\ & & {{}\Pi'} } } \phantom{mmmmm} \mbox{ \xymatrix @-4mm{ & & X^n\times \Pi \ar^{\text{id}\times p}[dd] \\ {}\Conf(X,n) \ar[r] \ar@/^.7pc/[urr] \ar@/_.7pc/[drr] & X^n \ar@{.>}_{\!\!\!G(\phi)}[ur] \ar@{.>}^{\!\!\!G(\psi)}[dr] & \\ & & X^n\times \Pi' } } \end{equation*} Now take the left of these two diagrams, where $`f$ and~$`q$ are rational maps defined on $\Conf(X,n)$, and notice that the projection $\mbox{id}\times p$ maps the closure $\overline{G(`f)}$ of the graph of~$`f$ onto the closure $\overline{G(`q)}$ of the graph of~$`q$. \end{proof} This surjection $\vartheta_n$ admits a more explicit description. For~$n\leqslant 3$, it is the identity map; otherwise, it can be written as a composition \begin{equation} \label{theta} \xymatrix @+3pt { X\pd<n>= W_{n-2}\ar[r]^<(0.3){\beta_{n-2}} & W_{n-3}\ar[r]^<(0.35){\beta_{n-3}} & {}\dots\ar[r]^{\beta_{3}} & W_2 \ar[r]^<(0.25){\beta_{2}} & W_1=X[n], } \end{equation} where $W_k \xrightarrow{`b_k} W_{k-1}$ is the blowup in~$W_{k-1}$ of (the disjoint union of the proper transforms under $`b_{k-1}\circ\dots\circ `b_2$ of) some strata $X(\S)$ of~$X[n]$; their encoding nests~$\S$ are characterized below. Favoring imprecision over repetitiveness, I will neglect to reiterate the ritual phrase that in the previous sentence appears in parentheses. Let $U\subset X[n]$ be the union of all strata $X(\S)$ such that the nest $\S$ contains two disjoint subsets of~$[n]$. The irreducible components of this codimension~2 reduced subscheme are $X(\S)$ for all nests $\S=\{S_1,S_2\}$ with $S_1\cap S_2=\varnothing$, which intersect transversally \cite[Theorem~3]{FM}. The map~$\vartheta_n$ is an iterated blowup of $X[n]$ along~$U$, but not all the strata contained in~$U$ are centers of a blowup $`b_k$. The components of the center of $`b_k$ are the strata $X(\S)$ such that $\S$ is the set of all essential blocks of a partition $`p\in\Ln$ with $`r(`p)=k$ and $`e(`p)>1$, which is always a nest. The transversality of the strata guarantees that, whenever the sequence~$`b_2,\dots,`b_{n-2}$ calls for two intersecting strata to be in the center of the same~$`b_k$, the previous stages will have made them disjoint. The sequence itself implies that, whenever $X(\S)\subset X(\S')$ are both to become centers, the smaller stratum is blown up before the larger one. Alternatively, the variety $W_k$ can be defined as the closure of $\Conf(X,n)$ in $$ X^n\times \prod_{`r(`p)\<k\text{ or }`e(`p)=1} \Bl_{`D^`p}X^n, $$ and an argument similar to Proposition~\ref{closure} shows that this is equivalent to the blowup description. \begin{examples} Here $X(S_1,\dots,S_k)=D(S_1)\cap\dots\cap D(S_k)$ refers to strata of~$X[n]$. The map $\vartheta_4$ is the blowup of 3 disjoint codimension-2 strata $X(12,34)$ and alike, for the nests obtained from the 3 partitions of shape $(2,2)$. The divisor $D^{12|34}\subset X\pd<4>$ is a $\P^1$-bundle over $X(12,34)$. For $n=5$, there are two maps in Eq.~(\ref{theta}). The first blows up 10 disjoint codimension-2 strata, like $X(123,45)$, corresponding to the partitions of shape $(3,2)$. The second blows up 15 disjoint codimension-2 strata, like $X(12,34)$, corresponding to $(2,2,1)$. For $n=6$, there are three stages according to the partitions $$ (4,2), \ \ \ (3,3); \quad\quad\quad (3,2,1), \ \ \ (2,2,2); \quad\quad\quad (2,2,1,1). $$ Here we encounter inclusions like $X(12,34,56)\subset X(12,34)$. Interestingly, the proper transform by $\vartheta_6$ of $X(12,34,56)$, which is the divisor $D^{12|34|56}$, is a bundle over $X(12,34,56)$ with fiber $\P^2$~blown up at three points. Proposition~\ref{small-bricks} generalizes this observation. \end{examples} The preimage in~$\Xn$ of a stratum of~$X[n]$ is $$ \vartheta_n^{-1}(X(\S))=\bigcup_{(T,`y)\in`h^{-1}(T(\S))} S_{(T,`y)}, $$ the union of all strata encoded by the leveled trees $(T,`y)$ with the same base tree $T(\S)$ and any legal assignment of levels to its interior vertices. The map~$\vartheta_n$ is thus strata-compatible. \begin{prop} The fibers of $\vartheta_n \colon \Xn \to X[n]$ are independent of~$X$ and even of its dimension. The fiber over a point in $X(\S)$ that is not in any smaller stratum is completely determined by the nest~$\S$. \end{prop} \begin{proof} The normal space~$N_\x$ at a point~$\x$ to $X(\S) \subset X[n]$ is independent of $\dim X$ (assumed positive): its dimension is equal to the cardinality of the nest~$\S$. The nest alone determines the iterated blowup of~$N_\x$ induced from Eq.~(\ref{theta}), and the preimage of the origin under it is isomorphic to $\vartheta_n^{-1}(\x)$. \end{proof} \section{Structure of the strata} \label{sec-strata} \def{`p_1}{{`p_1}} \def{`p_2}{{`p_2}} \def\mathcal B{\mathcal B} This section begins by discussing a family of linear subspace arrangements indexed by integer partitions; each of them leads to a projective variety that will be called a brick. Points of a brick correspond to polyscreens, and by presenting the strata of~$\Xn$ as bundles over~$X\pd<k>$ whose fibers are products of bricks, the polyscreen description of~$\Xn$ is established here. The configuration space $\Conf(\mathbb A^{\!1},n)$ is the complement to the braid arrangement of hyperplanes in $\mathbb A^{\!n}$, the motivating example for much of the theory of hyperplane arrangements~\cite{OrlikTerao}. The analogue for $\Am$, denoted by~$\bar\mathcal B^m_n$, is an arrangement of codimension~$m$ linear subspaces of~$(\Am)^n$. Its strata are various intersections of the large diagonals, so the partitions of~$[n]$ index them, for each $m \> 1$; in other words, the intersection lattice of~$\bar\mathcal B^m_n$ is isomorphic to the partition lattice~$\Ln$. These and all other subspace arrangements encountered in this section are $c$-plexifications of hyperplane arrangements~\cite{Bjorner}. This means practically that most information about~$\bar\mathcal B^m_n$ can be extracted from the braid arrangement~$\bar\mathcal B^1_n$. For any partition~$`p$ of~$[n]$, the images in the quotient $C^m_`p = (\Am)^n / `D^`p$ of those large diagonals that contain $`D^`p$ form an induced arrangement $\mathcal B^m_`p$. For $`p = \bot(\Ln)$, it is denoted by $\mathcal B^m_{n-1}$ (actual subscripts will be integers~$`n_i$); if~$m=1$, this is the Coxeter arrangement of type~$A_{n-1}$. For other partitions, $\mathcal B^m_`p$ is a product arrangement, as Lemma~\ref{product-arrangement} shows below. \def\mathcal A{\mathcal A} \def{\vphantom{1}}{{\vphantom{1}}} For two subspace arrangements $\mathcal A_i = \{K^i_1,\dots,K^i_{s_i}\}$ in $\Bbbk$-vector spaces~$V_i$, $i=1,2$, the product arrangement $\mathcal A_1 \times \mathcal A_2$ in $V_1 \oplus V_2$ is the collection of subspaces $\{K^1_1 \oplus V^{\vphantom{1}}_2, \dots, K^1_{s_1} \oplus V^{\vphantom{1}}_2, K^2_1 \oplus V^{\vphantom{1}}_1, \dots, K^2_{s_2} \oplus V^{\vphantom{1}}_1\}$. For each integer partition $`l = (`n_1,\dots,`n_r)$, define~$\mathcal B^m_`l$ as the product $\mathcal B^m_{`n_1} \times\dots\times \mathcal B^m_{`n_r}$. The intersection lattice of a product is the product of those of the factors; for~$\mathcal B^m_`l$ this gives the lattice $L_`l = L_{[`n_1+1]} \times\dots\times L_{[`n_r+1]}$. As an example, take for $`l$ the finest partition $(1,\dots,1)$ of~$r$, often denoted by $1^r$. Since~$\mathcal B^1_1$ is the arrangement $\{0\}$ in~$\Bbbk$, its $r$-th power $\mathcal B^1_{1^r}$ is the arrangement of coordinate hyperplanes in~$\Bbbk^r$. \begin{lem} \label{product-arrangement} Up to a change of coordinates $\mathcal B^m_`p \simeq \mathcal B^m_`l$, where $`l=`l(`p)$ is the essential shape of~$`p$. \end{lem} \begin{proof} Look at the equations of the large diagonals containing~$`D^`p$, that is, $`D^{ij} = \{(x_1,\dots,x_n) \in (\Am)^n \thinspace|\thinspace x_i = x_j\}$ for all pairs of $i$~and~$j$ belonging to the same block of~$`p$. Equations coming from different blocks of~$`p$ are independent of each other, leading to the product decomposition. \end{proof} The polydiagonal compactification $\Am\pd<n>$ is the maximal blowup of the arrangement~$\bar\mathcal B^m_n$, in the sense that all strata of~$\bar\mathcal B^m_n$ are blown up in the course of its construction. In the same fashion, all strata of the arrangement~$\mathcal B^m_`l$ can be blown up in the ascending order given by their dimensions. The first stage is always the blowup of the origin, creating the exceptional divisor $\P(C^m_`l) \simeq \P^{m|`l|-1}$, where $|`l|$ is the sum of all parts of~$`l$. The main objects of interest for this section are defined as follows. \begin{defn} For any integer partition~$`l$, a \newterm{brick} $M^m_`l$ is the proper transform of $\P(C^m_`l)$ in the maximal blowup of $\mathcal B^m_`l$. If~$`l$ has only one part, the brick~$M^m_`l$ is \newterm{simple}, otherwise it is \newterm{compound}. The \newterm{open} brick $\kern-1pt M^m_`l$ is the complement in $\P(C^m_`l)$ of the projectivization of~$\mathcal B^m_`l$. \end{defn} \begin{examples} The brick $M^m_1$ is just $\P^{m-1}$ (a single point if $m=1$). The bricks $M^m_2$ and $M^m_{1,1}$ are blowups of $\P^{2m-1}$; their centers are, respectively, three and two copies of $M^m_1$. The bricks $M^m_3$, $M^m_{2,1}$ and $M^m_{1,1,1}$ are 2-stage blowups of $\P^{3m-1}$; the lower intervals in $L_3$, $L_{2,1}$ and $L_{1,1,1}$ determine their centers, respectively: \begin{list}{}{} \item 7 copies of $M^m_1$, then 6 copies of $M^m_2$; \item 4 copies of $M^m_1$, then 3 copies of $M^m_{1,1}$ and 1 copy of $M^m_2$; \item 3 copies of $M^m_1$, then 3 copies of $M^m_{1,1}$. \end{list} For $M^1_3$, look again at Figures 2~and~3 on page~\pageref{fig-x4}. Similar pictures for $M^1_{2,1}$ and $M^1_{1,1,1}$ are in Figures 10~and~11. Comparison of these figures suggests that refining the indexing partition corresponds to omitting some subspaces from the arrangement. This is proved in general in Proposition~\ref{sub-and-super}. \begin{figure}[t] \label{fig-four-lines} \def\kern-1pt{\kern-1pt} \xy (16,-12)*{\mycaptionempty}; (0,-6); (0,18)**\dir{-}; (-14,0); (14,0)**\dir{-}; (11.4,-5.1); (-3.6,17.4)**\dir{-}; (-11.4,-5.1); (3.6,17.4)**\dir{-}; (0,0)*{\mydot}; (8,0)*{\mydot}; (0,12)*{\mydot}; (-8,0)*{\mydot}; (3.9,15.1)*{\scriptstyle 1\!2}; (9,-4)*{\scriptstyle 1\!3}; (-12.6,1.2)*{\scriptstyle 4\05}; (-1.4,-4.2)*{\scriptstyle 2\03}; (24,-4)="m", (8,0)+"m"="bot", (0,7)+"m"="12", (4,7)+"m"="13", (8,7)+"m"="23", (16,7)+"m"="45", (0,14)+"m"="123", (8,14)+"m"="1245", (12,14)+"m"="1345", (16,14)+"m"="2345", (8,21)+"m"="top", @={"top","123","12","bot","13","123","23","bot","45","2345","23", "2345","top","1245","12","1245","45","1345","13","1345","top"}, s0="prev" @@{;"prev";**@{-}="prev"}, @i; @={"top","1245","1345","2345","123","12","13","23","45","bot"}, @@{*{\mydot}}, "12"+(-1.8,-1)*{{}^{1\!2}}, "13"+(-1.5,-1)*{{}^{1\!3}}, "23"+(2,-1)*{{}^{2\03}}, "45"+(2,-1)*{{}^{4\05}}, (77,0)="m", "m"+(16,-12)*{\mycaptionempty}; "m"+(-14,0); "m"+(6,0)**\dir{-}; "m"+(0,-6); "m"+(0,18)**\dir{-}; "m"+(-11.4,-5.1); "m"+(3.6,17.4)**\dir{-}; "m"+(0,0)*{\mydot}; "m"+(0,12)*{\mydot}; "m"+(-8,0)*{\mydot}; "m"+(3.9,15.1)*{\scriptstyle 1\!2}; "m"+(-1.5,-4.2)*{\scriptstyle 3\04}; "m"+(-12.6,1.2)*{\scriptstyle 5\06}; (16,-4)+"m"="m", (8,0)+"m"="bot", (0,7)+"m"="12", (8,7)+"m"="34", (16,7)+"m"="56", (0,14)+"m"="1234", (8,14)+"m"="1256", (16,14)+"m"="3456", (8,21)+"m"="top", @i; @={"1234","12","1256","56","3456","34","1234"}, s0="prev" @@{;"prev";**@{-}="prev"}, @i; @={"1234","1256","3456"}, "top"; @@{**@{-}}, @i; @={"12","34","56"}, "bot"; @@{**@{-}}, @i; @={"top","1234","1256","3456","12","34","56","bot"}, @@{*{\mydot}}; "12"+(-1.8,-1)*{{}^{1\!2}}, "34"+(2,-1)*{{}^{3\04}}, "56"+(2,-1)*{{}^{5\06}}, \endxy \end{figure} Of special importance is the brick~$M^1_{1^r}$ that arises from the coordinate arrangement in~$\Bbbk^r$: blow up $r$~points in $\P^{r-1}$ in general position, then blow up the proper transforms of all lines spanned by pairs of these points, then blow up those of all planes spanned by triples, and so on. Thus $M^1_{1^r}$ is isomorphic to the space~$\Pi_r$ that Kapranov called the {\em permutahedral space} \cite[p.~105]{Kapranov:Chow}. It~is the compact projective toric variety whose encoding polytope is the permutahedron~$P_r$, usually defined as the convex hull of the set of $r!$ points in~$`R^r$ with coordinates $(`s^{-1}(1),\dots,`s^{-1}(r))$, for all $`s\in\Sym_r$. This polytope can also be obtained from the standard $(r-1)$-simplex by chopping off first all its vertices, then all that remains of its edges, then faces, and so on; this corresponds to the sequence of blowups producing~$\Pi_r$. In addition, this variety is the closure of a principal toric orbit in the complete flag variety and it has been extensively studied from various perspectives \cite{Atiyah,DolgachevLunts,GelfandSerganova,Procesi,Stanley:log-concave,% Stembridge:Eulerian,Stembridge:reps}. For each $m \> 1$, the brick $M^m_{1^r}$ is a toric variety because all strata of~$\mathcal B^m_{1^r}$, sitting in $\P^{rm-1}$, are $(\kt)^{rm}$-invariant. \end{examples} \begin{prop} \label{compound} Every open compound brick has the structure of a bundle \begin{equation} \label{decompose} \xymatrix @-3mm{ {}\kern-1pt M^1_{1^r} \ar[r] & \kern-1pt M^m_{\lambda} \ar[r] & \kern-1pt M^m_{\nu_1}\times\dots\times\kern-1pt M^m_{\nu_r},} \end{equation} where $`l$ is the integer partition $(`n_1,\dots,`n_r)$. \end{prop} \begin{proof} The complement to $\mathcal B^m_{`n_i}$ in $(\Am)^{`n_i}$ is $\Conf(\Am,`n_i+1)/\Am$, the configuration space of $`n_i+1$ distinct labeled points in~$\Am$ modulo translations. Since~$\kern-1pt M^m_`l$ is the complement in $\P(C^m_`l)$ to the projectivization of the arrangement $\mathcal B^m_`l = \mathcal B^m_{`n_1} \times\dots\times \mathcal B^m_{`n_r}$, it follows that \begin{equation} \label{orbit} \kern-1pt M^m_\lambda = \P(\kern-1pt C^m_\lambda), \quad\text{where}\quad \kern-1pt C^m_\lambda = \prod_{i=1}^{r} \BBigfactor{\Conf(\Am,\nu_i+1)}{\Am}, \end{equation} is the orbit space of the diagonal action of~$\kt$ on this product by dilations. Separate actions of~$\kt$ on each factor together give that of~$(\kt)^r$ on~$\kern-1pt C^m_`l$. Its total orbit space is isomorphic to the product of those coming from the factors, which is $\kern-1pt M^m_{`n_1} \times\dots\times \kern-1pt M^m_{`n_r}$. The orbit space $\kern-1pt M^m_`l$ maps into this product, with fiber $(\kt)^r/\kt \simeq \kern-1pt M^1_{1^r}$. \end{proof} The next two propositions follow from the general work of De Concini and Procesi \cite[pages~480--482]{DeCP}, but they can also be proved directly. \begin{prop} \label{affine} \begin{enumerate} \item \label{affa} The compactification $\Am\pd<n>$ is the product $\Am \times \L$, where~$\L$ is the total space of a line bundle over the simple brick~$M^m_{n-1}$. \item \label{affb} The simple brick~$M^m_{n-1}$ is a compactification of\/ $\Conf(\Am,n)/\text{\rm Aff}$, where $\text{\rm Aff}$ is the group of all affine transformations in~$\Am$. \end{enumerate} \end{prop} \begin{proof} (\ref{affa}) The direct factor~$\Am$ is the small diagonal $`D \subset (\Am)^n$. The essential shape of the bottom partition of $[n]$ is the integer $n-1$, thus by definition, there is a map $\psi \colon M^m_{n-1} \to P = \P((\Am)^n/`D)$. The bundle~$\L$ is the pullback by~$\psi$ of the tautologial line bundle over~$P$; since~$\psi$ is an iterated blowup, Lemma~\ref{blow-bundle} (formulated below) has to be used at each stage. (\ref{affb}) Affine transformations identify any nondegenerate configuration in~$\Am$ with a degenerate one in which all $n$~points collide at $0$, cancelling both the direct factor~$\Am$ and the fiber of the line bundle~$\L^n$. \end{proof} \begin{lem} \label{blow-bundle} Let~$V$ be a smooth subvariety of a smooth algebraic variety~$W$, let $h\colon F\to W$ be a vector bundle over~$W$, and~$E$ its restriction onto~$V$. Then $\Bl_E F$ is a vector bundle over $\Bl_V W$ isomorphic to the pullback of~$F$ by the blowup projection. \end{lem} \begin{proof} The normal bundle $\normal{E}{\!F}$ is the pullback $h^*\normal{V}{W}$. \end{proof} There are two differences between the construction of $\Am\pd<n>$ and that of the bricks: different arrangements to start with and projectivization; both are minor enough that some basic facts about bricks follow by the same arguments that apply to $\Am\pd<n>$. In turn, describing first the strata of the bricks provides a quick way of doing the same for~$\Xn$. \begin{prop} \label{brick-strata} For any integer partition~$`l$, fix a partition $`p$ of $[n]$ of essential shape~$`l$ and an isomorphism $[`p,\top] \simeq L_{`l}$. \begin{enumerate} \item \label{brick-divisors} For each partition~${`p_1}$, $`p < {`p_1} < \top$, there is a divisor~$E^{`p_1}$ in~$M^m_`l$. The union of these divisors is the complement\/ $M^m_`l \smallsetminus \kern-1pt M^m_`l$, and any set of them meets transversally. \item \label{brick-chains} An intersection $E^{`p_1}\cap\dots\cap E^{`p_k}$ is nonempty exactly when the partitions form a chain. Thus $M^m_`l$ is stratified by strata parametrized by all chains in~$L_`l$ that include neither its bottom nor its top. \item \label{brick-products} For any such chain $\{`p_1,\dots,`p_k\}$, the corresponding stratum of\/~$M^m_`l$ is isomorphic to $M^m_{`l_0} \times\dots\times M^m_{`l_k}$, where the integer partitions $`l_0,\dots,`l_k$ are determined by $L_{`l_i} \simeq [`p_i,`p_{i+1}]$, with $`p_0 = `p$ and\/ $`p_{k+1} = \top$. \end{enumerate} \end{prop} \begin{defn} A smooth subvariety~$V$ of a smooth algebraic variety~$W$ will be called \newterm{straight} if the normal bundle $N_V W$ is isomorphic to a direct sum of copies of a single line bundle. In this case, the exceptional divisor of the blowup $\Bl_V W$ is a trivial bundle. \end{defn} \begin{lem} \label{straight} \begin{enumerate} \item \label{P-in-P} For any two positive integers\/ $k$~and\/~$l$, any linear subvariety\/ $\P^k$ of\/~$\P^{k+l+1}$ is straight. {\rm (}\!Whence the term.{\rm )} \item \label{blow-straight} Let\/ $Z$~and\/~$V$ be smooth subvarieties of a smooth algebraic variety~$W$, such that either\/ $Z\cap V=\varnothing$ or $Z\subset V$. If\/~$V$ is straight in~$W$, then so is its proper transform~$\tilde{V}$ in $\tilde{W}=\Bl_Z W$. \end{enumerate} \end{lem} \begin{proof} Part (\ref{P-in-P}) follows directly from the definition. (\ref{blow-straight}) Nothing to be done when $V$~and~$Z$ are disjoint. When $Z\subset V$, denote by~$E$ the exceptional divisor of $\tilde{W}$, and by~$p$ the projection $\tilde{V}\to V$, then $$ \normal{\tilde{V}}{\tilde{W}} \simeq p^*\normal{V}{W}\otimes\O(-E)\big|_{\tilde{V}}, $$ and the claim follows. \end{proof} \begin{proof}[Proof of Proposition~\ref{brick-strata}] Similarly to Proposition~\ref{divisor}, the definition of~$M^m_`l$ implies parts (\ref{brick-divisors})~and~(\ref{brick-chains}). Part~(\ref{brick-products}) can be checked by induction on~$k$, where the inductive step follows by applying the case $k=1$. Thus, it is enough to show that each divisor~$E^`p$ is isomorphic to $M^m_{`l_0} \times M^m_{`l_1}$, where $L_{`l_0} \simeq [`p,{`p_1}]$ and $L_{`l_1} \simeq [{`p_1},\top]$. The argument is based on Lemmas \ref{blow-bundle}~and~\ref{straight}. Every partition from $[`p,\top]$ belongs to one of the following six groups: \begin{equation*} \begin{align*} \text{(i)} \ &\ \{`p\}, &&& \text{(ii)} \ &\ \{`p' \thinspace|\thinspace `p<`p'<{`p_1}\}, &&\\ \text{(iii)} \ &\ \{{`p_1}\}, &&& \text{(iv)} \ &\ \{`p' \thinspace|\thinspace {`p_1}<`p'<\top\}, &&\\ \text{(v)} \ &\ \{\top\}, &&& \text{(vi)} \ &\ \text{incomparable with }~{`p_1}. \end{align*} \end{equation*} The proof will be completed by studying the impact of blowups corresponding to partitions in each group on the stratum $`D^{`p_1} / `D^`p$ of the arrangement $\mathcal B^m_`l$. Before the blowups, the arrangements induced in $`D^{`p_1} / `D^`p$ and $(\Am)^n / `D^{`p_1}$ are isomorphic respectively to $\mathcal B^m_{`l_0}$ and $\mathcal B^m_{`l_1}$. First group, first stage. The exceptional divisor $\P(C^m_`l)$ of the first stage has a straight subvariety $\P( `D^{`p_1} / `D^`p) \simeq \P(C^m_{`l_0})$ with the projectivization of~$\mathcal B^m_{`l_0}$ in it, and with the arrangement $\mathcal B^m_{`l_1}$ in each normal space to it (Lemma~\ref{blow-bundle}). Lemmas \ref{blow-bundle}~and~\ref{straight} also apply at the subsequent stages, pulling back arrangements inside normal spaces and preserving the straightness of blowup centers. Group~(vi) blowups are irrelevant for the divisor~$E^`p$ at all stages, and no blowup corresponds to~$\top$. Group~(ii) blowups turn $\P(C^m_{`l_0})$ into $M^m_{`l_0}$. Then the group~(iii) blowup makes it into a divisor isomorphic to $M^m_{`l_0} \times \P(C^m_{`l_1})$. The second factor inherits the projectivization of~$\mathcal B^m_{`l_1}$, and blowups of the remaining group~(iv) transform this divisor into $E^{`p_1} \simeq M^m_{`l_0} \times M^m_{`l_1}$. \end{proof} \begin{lem} \label{divisor-bundle} Each divisor $D^`p$ of\/ $\Xn$ is isomorphic to a bundle over $X\pd<`r(`p)>$ with fiber~$M^m_{`l(`p)}$. In addition, this bundle is trivial if\/~$X=\Am$. \end{lem} \begin{proof} Corollary~\ref{middleX} gives $Y^`p_{r-1} \simeq X\pd<r>$, where $r =`r(`p)$. By~Lemma~\ref{blow-bundle}, the arrangements~$\mathcal B^m_`p$ transform isomorphically from the normal spaces to $`D^`p$ in $\Xn$ into the normal spaces to $Y^`p_{r-1}$ in $Y^{\vphantom{`p}}_{r-1}$. At the next stage, $Y^`p_r$ is a bundle over $X\pd<r>$ with fibers isomorphic to $P = \P((\Am)^n/`D)$. The relevant blowup centers of the subsequent stages are its subbundles; their fibers form in every fiber of $Y^`p_r$ an arrangement isomorphic to the projectivization of $\mathcal B^m_`p$ in~$P$. Thus in the end, the fibers of $Y^`p_r$ transform into $M^m_{`l(`p)}$. If in addition $X=\Am$, a repeated application of Lemma~\ref{straight} shows that $Y^`p_{r-1}$ is straight in $Y^`p$, so $Y^`p_r = \Am\pd<r> \times P$ and the result follows. \end{proof} \begin{prop} \label{tower} Let\/~$`g = \{`p_1,\dots,`p_k\}$ be a chain of partitions of\/~$[n]$ and let\/ $\{`l_0,\dots,`l_k\}$ be its associated sequence of integer partitions (Section~\ref{sec-combinat}). \begin{enumerate} \item The stratum~$S_`g$ of\/~$\Xn$ is a bundle over $X\pd<`l_0>$ with fiber isomorphic to $M^m_{`l_1} \times\dots\times M^m_{`l_k}$. \item Consequently, the complement in~$S_`g$ to the union of smaller strata, the open stratum $\kern-1pt S_`g$, is a bundle over $\Conf(X,`l_0)$ with fiber isomorphic to $\kern-1pt M^m_{`l_1} \times\dots\times \kern-1pt M^m_{`l_k}$. \end{enumerate} \end{prop} \begin{proof} Put together Lemma~\ref{divisor-bundle} and Proposition~\ref{brick-strata}. \end{proof} By this proposition, a~point in a stratum~$\kern-1pt S_`g$ is given by a~configuration of $r$~distinct points in~$X$ (where the collision occurs) and a sequence consisting of one point in each open brick $\kern-1pt M^m_{`l_i}$. Equations (\ref{decompose})~and~(\ref{orbit}) in Proposition~\ref{compound} show that such points can be represented by suitable polyscreens: points in each constituent open simple brick are Fulton--MacPherson screens, and points in $\kern-1pt M^1_{1^r}$ are $r$-tuples of scale factors. Thus, points of~$\Xn$ indeed have the geometric description explained in Section~\ref{sec-screens}. \begin{prop} \label{small-bricks} The compound brick\/ $M^m_{1^r}$ has the structure of a bundle \begin{equation} \label{only-ones} \xymatrix @-3mm{ \Pi_{r} \ar[r] & M^m_{1^r} \ar[r] & (M^m_1)^r. } \end{equation} \end{prop} \begin{proof} Fix a partition~$`p$ of $[2r]$ into two-element blocks and let the nest $\S$ be the set $\{`b_1,\dots,`b_r\}$ of blocks of~$`p$. The map $\vartheta_{2r} \colon \Am\pd<2r> \to \Am[2r]$ takes the divisor~$D^`p$ of $\Am\pd<2r>$ into the stratum $\Am(\S)$ of $\Am[2r]$. The divisor is isomorphic to $\Am\pd<r> \times M^m_{1^r}$ by Lemma~\ref{divisor-bundle} and the stratum is isomorphic to $\Am[r] \times (\P^{m-1})^r$. Since $M^m_1 \simeq \P^{m-1}$, it~follows that $M^m_{1^r}$ maps to $(M^m_1)^r$. Tracing $\vartheta_{2r}^{-1}$ stage by stage, first transform the factor $\Am[r]$ into $\Am\pd<r>$; then at stage~$r$ blow up the proper transform of $\Am(\S)$, turning the second factor into a $\P^{r-1}$-bundle over $(M^m_1)^r$. Since $\Am[n] \smallsetminus \Conf(\Am,n)$ is a normal crossing divisor, the divisors $D(`b_i)$ induce in each fiber $\P^{r-1}$ the projectivized coordinate hyperplane arrangement. All of its strata are blown up at the subsequent stages, turning $\P^{r-1}$ into $M^1_{1^r} \simeq \Pi_r$. \end{proof} \begin{figure}[t] \label{levels-split} \xy (-14,-17)*{\ }; (-14,9)*{\ }; (80,-12)*{\mycaption{5in}{One level splits into two}}, (0,0)="m" *{\mydot}; @i @={(0,3)+"m",(1.5,-3)+"m",(-1.5,-3)+"m"}, "m"; @@{**@{-}}, (6,0)+"m"="m" *{\mydot}; @i @={(0,3)+"m",(1.5,-3)+"m",(-1.5,-3)+"m"}, "m"; @@{**@{-}}, (6,0)+"m"="m" *{\mydot}; @i @={(0,3)+"m",(1.5,-3)+"m",(-1.5,-3)+"m"}, "m"; @@{**@{-}}, (6,0)+"m"="m" *{\mydot}; @i @={(0,3)+"m",(1.5,-3)+"m",(-1.5,-3)+"m"}, "m"; @@{**@{-}}, (6,0)+"m"="m" *{\mydot}; @i @={(0,3)+"m",(1.5,-3)+"m",(-1.5,-3)+"m"}, "m"; @@{**@{-}}, (6,0)+"m"="m" *{\mydot}; @i @={(0,3)+"m",(1.5,-3)+"m",(-1.5,-3)+"m"}, "m"; @@{**@{-}}, (6,0)+"m"="m" *{\mydot}; @i @={(0,3)+"m",(1.5,-3)+"m",(-1.5,-3)+"m"}, "m"; @@{**@{-}}; (-2,3.2); (38,3.2)**\dir{.}; (-2,-3.2); (38,-3.2)**\dir{.}; (18,5)*\txt{\scriptsize $\cdots$ upper levels $\cdots$}; (18,-5)*\txt{\scriptsize $\cdots$ deeper levels $\cdots$}; (60,0)="m", @i @={(0,4)+"m",(1.5,-4)+"m",(-1.5,-4)+"m"}, "m"-(0,2); @@{**@{-}}; *{\mydot}; (6,0)+"m"="m" @i @={(0,4)+"m",(1.5,-4)+"m",(-1.5,-4)+"m"}, "m"+(0,2); @@{**@{-}}; *{\mydot}; (6,0)+"m"="m" @i @={(0,4)+"m",(1.5,-4)+"m",(-1.5,-4)+"m"}, "m"-(0,2); @@{**@{-}}; *{\mydot}; (6,0)+"m"="m" @i @={(0,4)+"m",(1.5,-4)+"m",(-1.5,-4)+"m"}, "m"+(0,2); @@{**@{-}}; *{\mydot}; (6,0)+"m"="m" @i @={(0,4)+"m",(1.5,-4)+"m",(-1.5,-4)+"m"}, "m"+(0,2); @@{**@{-}}; *{\mydot}; (6,0)+"m"="m" @i @={(0,4)+"m",(1.5,-4)+"m",(-1.5,-4)+"m"}, "m"-(0,2); @@{**@{-}}; *{\mydot}; (6,0)+"m"="m" @i @={(0,4)+"m",(1.5,-4)+"m",(-1.5,-4)+"m"}, "m"-(0,2); @@{**@{-}}; *{\mydot}; "m"+(2,4.2); "m"+(-38,4.2)**\dir{.}; "m"+(2,-4.2); "m"+(-38,-4.2)**\dir{.}; "m"+(-18,6)*\txt{\scriptsize $\cdots$ upper levels $\cdots$}; "m"+(-18,-6)*\txt{\scriptsize $\cdots$ deeper levels $\cdots$}; \endxy \end{figure} The fiber~$\Pi_r$ in Eq.~(\ref{only-ones}) stores scale factors; points in its open part $\kern-1pt M^1_{1^r}$ are generic and each is a part of one polyscreen. The divisor $E_r = \Pi_r \smallsetminus \kern-1pt M^1_{1^r}$ has components isomorphic to $\Pi_{s} \times \Pi_{r-s}$, whose points represent degenerations with $s$~scale factors tending to zero, and therefore polyscreens that split into two: $s$~screens form a~new level. For example, the left leveled tree in Figure~12 may degenerate into the right one, corresponding to a divisor $\Pi_4 \times \Pi_3 \subset \Pi_7$. The new levels may of course split further; intersections of components of~$E_r$ give a stratification of~$\Pi_r$, and each stratum is a product of a number of smaller permutahedral varieties. This corresponds to the well-known fact that all faces of the permutahedron~$P_r$ are products of lower-dimensional permutahedra~\cite{BilleraSarangarajan}. Other compound bricks, that is, $M^m_`l$ for those integer partitions~$`l$ that have parts greater than~$1$, do not admit decompositions similar to Eq.~(\ref{only-ones}), but each of them is a blowup of~$M^m_{1^r}$ for $r = |`l|$. Let~$`L_r$ be the set of all partitions of an integer~$r$ partially ordered by refinement: $(5,3) < (4,2,1,1)$ in~$`L_8$ because $5 = 4+1$ and $3 = 2+1$. It turns out that the set of bricks $\{M^m_`l \thinspace|\thinspace `l \in `L_r\}$ has a compatible (reverse) \inquotes{blowing-up} partial order. \begin{prop} \label{sub-and-super} Suppose that\/ $`l,`l' \in `L_r$ and $`l <`l'$. \begin{enumerate} \item \label{sublattice} The lattice~$L_{`l'}$ contains a sublattice isomorphic to~$L_{`l}$. \item \label{subarrangement} The subarrangement of~$\mathcal B^m_{`l'}$ formed by the subspaces that correspond to this sublattice is~$\mathcal B^m_`l$, up to coordinate change. \item \label{superbrick} The brick~$M^m_{`l'}$ is an iterated blowup of\/~$M^m_{`l}$. \end{enumerate} \end{prop} \begin{proof} (\ref{sublattice}) It is enough to show this for $`l' = (r-1)$ and $`l = (s-1,r-s)$. The required sublattice of~$L_{[r]}$ is generated by the union $[{`p_1},\top] \cup [{`p_2},\top]$, where the only essential block of ${`p_1}$ (${`p_2}$) is $\{ k \thinspace|\thinspace k \< s \}$ (resp.~$\{ k \thinspace|\thinspace k \> s \}$). (\ref{subarrangement}) It is enough to consider the same $`l$~and~$`l'$ as in~(\ref{sublattice}) and then write explicitly the equations for the large diagonals. (\ref{superbrick}) The two lattices $L_`l \subset L_{`l'}$ determine the sequences of blowups of $\P^{rm-1}$ creating $M^m_`l$ and $M^m_{`l'}$. It suffices to show that the blowups making~$M^m_{`l'}$ can be rearranged, without changing the outcome (up to an isomorphism), into a different sequence so that an intermediate stage is $M^m_{`l}$. This situation is quite similar to the consideration of $\vartheta_n \colon \Xn\to X[n]$ in Section~\ref{sec-closure}, and similar is the solution. \end{proof} \section{Isotropy of the permutation action} \label{sec-isotropy} Assume that the ground field~$\Bbbk$ is of characteristic~$0$. Reading carefully into Fulton and MacPherson's proof of the solvability of the isotropy subgroups of~$\Sym_n$ acting on~$X[n]$, one soon realizes that every point where the isotropy subgroup fails to be abelian lies in a stratum whose encoding nest contains a pair of disjoint subsets of~$[n]$. Exactly these strata are blown up by $\vartheta_n\colon\Xn\to X[n]$, and this observation raises hopes that are not false. \begin{thm} \label{abelian} If\/~$X$ is a smooth algebraic variety over a field\/~$\Bbbk$ of characteristic~$0$, then all isotropy subgroups of\/~$\Sym_n$ acting on~$\Xn$ by permutations of labels are abelian. \end{thm} \begin{proof} First, reduce to the case of all~$n$ points colliding at the same point in~$X$. Suppose a collision~$\x$ occurs at $p_1,\dots,p_r\in X$. If it could be studied near each~$p_i$ independently of the other points, as for $X[n]$, the isotropy subgroup would have been $G^{p_1}\times\dots\times G^{p_r}$, where $G^{p_i}$ is the isotropy subgroup of the collision near~$p_i$. It would have corresponded to~$r$ independent sequences of colored screens and reduced the proof to the case of one collision point, but this does not suit~$\Xn$. Fortunately, interdependencies among the corresponding levels in those $r$~sequences only put more restrictions on a permutation aspiring to fix~$\x$. It means that the isotropy subgroup will be a subgroup of the product above, which still does the trick. Pick a $k$-chain $`g\ni\bot$ and a coherent sequence of colored screens $\CS^j(\x)$ for~$`g$. A~permutation $`s\in\Sym_n$ fixes~$\x\in\kern-1pt S_`g$ if and only if it fixes all $\CS^j(\x)$. A~colored screen is fixed by~$`s$ exactly when these two conditions are fulfilled: \begin{enumerate} \renewcommand\theenumi{\arabic{enumi}} \renewcommand\labelenumi{(F\theenumi)} \item it is fixed modulo colors; \item any two points of the same color go to two points of the same color, not necessarily the original one. \end{enumerate} Let~$G$ be the isotropy subgroup at~$\x$. A permutation $`s\in G$ satisfies~(F1) for $\CS^j(\x)$, therefore it induces the scaling of $T_pX$ underlying $\CS^j(\x)$ by a scale factor $f_j(`s)\in\kt$. The map $f_j\colon G\to\kt$ is a group homomorphism, thus there is a group homomorphism $(f_1,\dots,f_k)= f\colon G\to (\kt)^k$, and to show that it is injective suffices to complete the theorem. Take $`s\in\ker f$, then~$`s$ does not move points in any of the colored screens $\CS^j(\x)$, $j=1,\dots,k$. By coherence, every color in $\CS^j(\x)$ is a point in $\CS^{j-1}(\x)$, since both are but blocks of the partition $`p_j\in`g$. Thus, $`s$ cannot change colors either, in any $\CS^j(\x)$ for $j=2,\dots,k$, and colors in $\CS^1(\x)$ stay unchanged because there is only one such. This shows that~$`s$ does not move anything at all, and there is only one such permutation: if $`s\neq\operatorname{id}$ and $`s(a)=b$, then~$`s$ must induce nontrivial scaling on $\CS^l(\x)$, where~$l$ is the maximal index~$j$ for which $a$~and~$b$ are in the same block of $`p_j\in`g$. \end{proof} \begin{rem} This version of the original proof is one substantially simplified with a key idea due to Jean--Luc Brylinski. \end{rem}
\section{Introduction} The evolution of stars is considerably complicated by the presence of a close companion, which may lead to either extreme mass loss, to strong mass accretion, or even to the merging of both stars. An overview of the evolutionary possibilities in massive close binaries is given by Podsiadlowski et al. (1992), while evolutionary models for parts of the --- unfavorably large --- initial parameters space of binaries have been computed, e.g., by Paczinsky (1967), Kippenhahn (1969), de Loore \& De Greve (1992), Pols (1994); see also Vanbeveren (1998ab, and references therein). On the other hand, close binaries provide unique possibilities to test and constrain uncertainties inherent in the evolution of stars in general. An example is given by Ergma \&\ van den Heuvel (1998), who showed that massive close binary systems containing a compact stellar remnant (neutron star or black hole) can constrain the initial mass limit for black hole formation $M_{\rm BH}$. In the present paper, we want to pursue this idea in a quantitative way. In principle, the problem is simple: a valid progenitor model for a system containing a neutron star yields the initial mass of the neutron star progenitor and thus a lower limit to $M_{\rm BH}$, and systems containing a black hole yield an upper limit to $M_{\rm BH}$. This procedure is not hopeless, even though the calculations of the progenitor evolution of the observed systems can not provide the information whether a model component in its final stage evolves into a neutron star or a black hole: The observed orbital period together with the properties of the normal star in the system may allow only for a very limited range of initial masses for the progenitor of the compact component. On the other hand, all the uncertainties involved with the theoretical description of mass transfer in massive close binaries enter the problem and increase the error bar on the progenitor masses. Note that $M_{\rm BH}$ is not a well defined quantity in binary systems; i.e., whether a star in a binary system forms a black hole or a neutron star depends not only on its initial mass but also on its evolutionary history (see below). We assume implicitly in this paper that $M_{\rm BH}$ is in fact well defined for single stars. This is also not guaranteed. E.g., the initial rotation rate may be a parameter to be considered in addition to the initial mass (cf., Heger et al. 1999). Ergma \& van den Heuvel (1998) even argue from the properties of massive close binaries with compact companions that in single stars, $M_{\rm BH}$ must be depending on a second parameter. However, even though we can not show the contrary, we attempt here to disprove their main argument for this. Two criteria may help to chose those observed systems which are best suited to constrain $M_{\rm BH}$. First, if one would pick systems for which one could assume that the system progenitor evolution proceeded more or less conservatively, i.e. that most of the system initial mass remained in the system, one could avoid the huge uncertainties inherent in theories of mass outflow from the system, either through the second Lagrangian point or in the course of a common envelope evolution (Podsiadlowski et al. 1992). Since strong mass outflow efficiently removes angular momentum and thus results in short periods, one should avoid the systems with the shortest periods. Second, it would be most efficient to investigate systems of which one can hope that the compact star's progenitor mass is close to $M_{\rm BH}$. For systems containing a neutron star, this means one should look for the most massive systems. The massive X-ray binary Wray 977/GX~301-2 (4U 1223-62) fulfills both criteria. It contains an X-ray pulsar (GX~301-2) and the B supergiant Wray 977 (BP Cru). The latest and most reliable determination of the stellar parameters of Wray~977 has been performed by Kaper \&\ Najarro (1999). They found log ${\mathrm T_{eff}/[K] = 4.23...4.30}$ and a radius of $R = 60...70 \rso$, with preferred values of 4.23 and 62 $\rso$. A radius of 62 $\rso$ combined with the empirical mass function of Wray~977/GX 301-2 (Kaper et al. 1995) and the absence of eclipses implies a lower mass limit for the B star of $M=39...40 \mso$. An independent lower mass limit for Wray~977 of $\sim 40\mso$ has been derived spectroscopically from the velocity amplitude by Kaper \&\ Najarro (1999). Constraints on the properties of Wray 977 obtained by Koh et al. (1997) agree with these values. The observed period of the binary is 44.15 days. Ergma \&\ van den Heuvel (1998) concluded that the neutron star progenitor mass was initially larger than the present mass of the B~star. Here, we present massive close binary models from zero age until beyond the death of the primary component (defined here as the initially more massive star in the system), which constrain the progenitor evolution of Wray 977/GX~301-2. After describing our computational method in Section~2, we present our best model for Wray 977/GX~301-2's progenitor evolution in Section~3, which results in an initial mass for GX~301-2 of only 26$\mso$. In Section~4, we discuss the evolution of the stellar, the helium core, and the CO-core masses of primaries in massive binary system on the basis of a new grid of computed systems. In Section~5, our grid of binary models is used to derive the transformation of the initial mass limit for neutron star formation obtained from binaries to the single star case. In Section~6 we analyze the evolution of the chemical structure of massive primaries and implications for Type~Ib/c supernovae. In Section~7 we summarize our results to a global picture of massive close binary evolution and a scenario for the formation of black hole and neutron star binaries. \section{Computational methods} We computed the evolution of massive close binary systems using a computer code generated by Braun (1997) on the basis of an implicit hydrodynamic stellar evolution code for single stars (cf. Langer 1991, 1998). It invokes the simultaneous evolution of the two stellar components of a binary and computes mass transfer within the Roche approximation (Kopal 1978). The entropy of the accreted material is assumed to be equal to that of the surface of the secondary star, where gravitational energy release due to mass accretion is treated as in Neo et al. (1977); see also Braun \& Langer (1995). Even though mass loss due to stellar winds is included for both components (see below) --- with a corresponding angular momentum loss according to Brookshaw \& Tavani (1993) --- the present calculations deal only with contact-free evolutionary stages, and therefore no other source of mass outflow from the system is included. Our models can thus be called quasi-conservative. We use standard stellar wind mass loss rates, i.e., the rate of Nieuwenhuijzen \& de Jager (1990), except for hot stars. For OB stars ($\teff > 15\, 000\,$K), we use the theoretical radiation driven wind models of Kudritzki et al. (1989) with wind parameters $k=0.085$, $\alpha = 0.657$, $\delta = 0.095$, and $\beta =1$ (Pauldrach et al. 1994). I.e., the dependence of the mass loss rate on the luminosity, effective temperature, mass, and surface hydrogen mass fraction is taken into account. For hydrogen-poor stars, i.e. stars with a surface hydrogen mass fraction $X_{\rm s} < 0.4$, we used a relation based on the empirical mass loss rates of Wolf-Rayet stars derived by Hamann et al. (1995) for massive stars $\llso \ge 4.5$ and by Hamann et al. (1982) for helium stars in the range $3.5 \le \llso \le 4.5$, i.e. \[ \log(-\dot M_{\mathrm WR}/(\msoy))= \] \begin{equation} \hfill \left\{ \begin{array}{ll} \displaystyle -11.95+1.5\llso-2.85 X_{\rm s}&\hbox{\ for\ } \llso\ge 4.5\\ \rule{0pt}{5mm}-35.8+6.8\llso&\hbox{\ for\ } \llso<4.5\\ \end{array} \right. \end{equation} Note that for core helium burning helium stars, this equation gives mass loss rates very close to that proposed by Langer (1989). To account for recent revisions of empirical Wolf-Rayet mass loss rates by Hamann \& Koesterke (1998), who suggested that previously derived values for massive Wolf-Rayet stars may overestimate the mass loss by a factor of 2...3, we also computed sequences where this rate has been multiplied by a factor 0.5, i.e., \[ \log(-\dot M_{\mathrm WR}/(\msoy))= \] \begin{equation} \hfill \left\{ \begin{array}{ll} \displaystyle -12.25+1.5\llso-2.85 X_{\rm s}&\hbox{\ for\ } \llso\ge 4.45\\ \rule{0pt}{5mm}-35.8+6.8\llso&\hbox{\ for\ } \llso<4.45\\ \end{array} \right. \end{equation} Convection and semiconvection have been treated according to Langer et al. (1983), mostly using a semiconvective efficiency parameter of $\alpha_{\mathrm{sc}} = 0.01$ (cf. also Braun \& Langer 1995). Time-dependent thermohaline mixing is followed in a diffusion scheme according to Wellstein et al. (1999; cf. also Braun 1997), which is based on an analysis of Kippenhahn et al. (1980). Opacities are taken from Iglesias \& Rogers (1996). Changes in the chemical composition are computed using a nuclear network including the pp-chains, the CNO-tri-cycle, and the major helium-, carbon and oxygen burning reactions. For the $^{12}$C($\alpha$,$\gamma$)$^{16}$O nuclear reaction rate we followed Weaver \& Woosley (1993) and used a value of 1.7~times that of Caughlan \& Fowler (1988). Further details about the computer program and input physics can be found in Langer (1998) and references therein. All models in this work use an approximately solar initial chemical composition with a mass fraction of all elements heavier than helium (``metals'') of $Z=0.02$. The mass fractions of hydrogen and helium are set to $X=0.7$ and $Y=1-X-Z=0.28$, respectively. The abundance ratios of the isotopes for a given element are chosen to have the solar system meteoritic abundance ratios according to Grevesse \& Noels (1993). \section{The smallest possible initial mass of GX 301-2} The derivation of initial masses of neutron star progenitors in binaries can only yield lower limits to $M_\mathrm{BH}$. These limits are immediately valid for single stars, since the stronger mass loss in primaries of close binaries compared to that in single stars can cause more massive primaries than single stars to form neutron stars, but not vice versa. As due to the uncertainties in the binary evolution models one always ends up with a possible range of initial masses for the neutron star progenitor in a given system, only the smallest mass in this range leads to a stringent constraint on $M_\mathrm{BH}$. Therefore, we attempt to find the lowest possible initial mass for the progenitor of the neutron star GX~301-2. \subsection{Restricting the initial system parameters} \begin{figure*}[ht] \begin{centering} \epsfxsize=0.8\hsize \epsffile{8742f1.ps} \caption{Evolution of the components of our progenitor model for Wray~977/GX~301-2 (model No.~8; cf. Table~\ref{fig:systems}) in the HR diagram. The dashed line corresponds to the evolution of the primary from the zero age main sequence to its supernova explosion. The solid line corresponds to the evolution of the secondary from the zero age main sequence until the onset of reverse mass transfer onto the neutron star (solid line); the dotted continuation of the solid line designates the secondary's evolution up to its supernova explosion (marked by a asterisk) if the neutron star companion were absent. The end of the solid line corresponds to a central helium mass fraction of the secondary of $Y_{\rm c} = 0.14$. Beginning and end of the various mass transfer phases are marked as follows. 1: begin of Case~A, 2: end of Case~A, 3: begin of Case~AB, 4: end of Case~AB. A/a designates core hydrogen exhaustion for the primary/secondary, B the end of primary's core helium burning. The position of both components at the time of the primary's supernova explosion is marked by a diamond and labeled `SN1'. The position of Wray~977 according to Kaper \& Najarro (1999) is indicated (square) together with the error bars. } \label{fig:hrd} \end{centering} \end{figure*} Since the neutron star progenitor evolved first into a supernova, the B~star Wray~977 is very likely the initially less massive star in the binary system --- we thus designate it here as the secondary component. Note that even though a reversal of the supernova order --- i.e., the secondary component exploding earlier than the primary --- is possible (e.g., Pols 1994, or System No.~10 in Section~4 below), this assumption would lead to larger neutron star progenitor masses than the assumption of the normal supernova order. If we define, as usual, $\beta$ as the fraction of the mass transfered from the primary to the secondary which remains in the binary system, i.e. the fraction $1-\beta$ is ejected from the system, one can easily see that $\beta\rightarrow 1$ (i.e. conservative evolution) leads to the smallest possible initial masses for the neutron star progenitor. E.g., assuming $\beta =0$ implies that the progenitor mass of GX~301-2 must be larger than Wray~977's present mass, i.e., $M_{\rm 1,i} \gtrsim 40\mso$ (cf. Ergma \& van den Heuvel 1998), while every solar mass which is successfully transfered form the neutron star progenitor to Wray~977 allows the initial mass of GX~301-2 to have been roughly one solar mass smaller. Furthermore, smaller neutron star progenitor masses can be achieved the closer the initial mass ratio $q$ is to one. Keeping the total mass $M$ in the system constant up to the first supernova explosion ($\beta =1$), the smallest possible initial mass of the neutron star progenitor is $M/2$, which corresponds to $q=1$. Finally, we need to know the initial period of our most constraining binary model. There are two reasons which made us prefer a Case~A evolution (mass transfer during core hydrogen burning). First, simple estimates --- which are possible and meaningful for conservative systems --- show that initial periods corresponding to Case~A result in final periods which are in the range of that found in Wray~977/GX~301-2 (Pols 1994). Second, Case~A mass transfer yields smaller values for $M_{\rm BH}$ in single stars than Case~B or~C \footnote{as Kippenhahn \& Weigert (1967) and Podsiadlowski (1992), we define the Cases~A, B, and C evolution corresponding to mass transfer during core hydrogen burning, after core hydrogen burning but before core helium exhaustion, and after core helium exhaustion, respectively} (as will be shown in detail below), and as we are looking for the lower limit on $M_{\rm BH}$ we thus need to consider Case~A. \subsection{A progenitor model for Wray~977/GX~301-2} The binary model (System No.~8; cf. Table~1) with initial parameters selected in this way has a primary star initial mass of 26$\mso$, a secondary initial mass of 25$\mso$, and an initial period of 3.5 days. The evolution of both components in the HR diagram is displayed in Figure~\ref{fig:hrd}. It proceeds through Case~A mass transfer, for which a rapid and a slow phase can be distinguished and which is followed by a Case~AB mass transfer after the core hydrogen exhaustion of the primary (cf. Fig.~\ref{fig:m_trans}). The secondary evolves to a maximum mass of 42.4$\mso$, which is reduced by stellar winds later on. After the Case~AB mass transfer, the primary is a helium star of 6.4$\mso$ (i.e., a Wolf-Rayet star) which evolves to a final mass of 3.2$\mso$ due to wind mass loss (cf. Fig.~\ref{fig:m26conv}). Note that the initial helium star mass of this 26$\mso$ primary is smaller than the corresponding mass in a 25$\mso$ primary undergoing Case~B mass transfer (System No.~9 in Table~1 below), since in Case~A primaries the hydrogen burning convective core mass is reduced due to the mass transfer (cf. Fig.~\ref{fig:m26conv}). Our System No.~8 does not only represent the academic case which yields the minimum initial progenitor mass of the neutron star GX~301-2. It also fulfills all empirical constraints imposed by Wray~977/GX~301-2 (cf. Section~1, and see Fig.~\ref{fig:hrd}) and is thus a viable progenitor model for this system. The mass of the B~star at the time of the supernova explosion of the primary is 40.5$\mso$, and the final period of the system of 46.22~d before and 47.3~d after the supernova explosion --- without considering a supernova induced kick on the neutron star --- is in good agreement with the observed period. Note that the possibility of supernova kicks render conclusions based on the observed period as difficult since the observed eccentricity of $e=0.47$ is rather large; it only constrains the period of the spherical orbit before the supernova explosion approximately to the range 15$\,$d...59$\,$d. While the luminosity of the mass gainer in our model~No.8 ($\llso \simeq 5.7$) is within the observational error bar, it is slightly more luminous than the value preferred by Kaper \&\ Najarro (1999) ($\llso \simeq 5.5$). To analyze the uncertainties of the post-main sequence luminosity in our models, we have computed several $25 + 24\mso$ Case~A systems with different assumptions for the semiconvective efficiency parameter $\alpha_{\mathrm sc}$. This parameter, which controls the so called rejuvenation process in the accreting main sequence star (cf. Hellings 1983, Braun \& Langer 1995), has no influence on the evolution of the mass transfer, the stellar masses or the binary period. However, it does affect the evolutionary track of the secondary after the mass transfer. Most important, it determines the temperature and radius evolution after core hydrogen exhaustion, and to a smaller extent it influences the stellar luminosity during this phase. From Figure~\ref{fig:m25hrd} we see that smaller values of $\alpha_{\mathrm sc}$ lead to smaller post-main sequence luminosities. While the semiconvection parameter has no relevance for our discussion of the critical mass limit for neutron star formation, it is important for the probability to find systems like Wray~977/GX~301-2 --- which is much higher for lower $\alpha_{\mathrm sc}$ --- and for its future evolution (cf. Fig.~\ref{fig:m25hrd}). Note that System No.~8 has been computed with an efficiency parameter for semiconvection of $\alpha_{\mathrm sc} = 0.02$ (see Table~\ref{fig:systems}). \begin{figure*}[t] \begin{centering} \epsfxsize=0.8\hsize \epsffile{8742f2.ps} \caption{Evolution of the internal structure of the primary component of System No.~8 --- i.e., our progenitor model for GX~301-2 --- as function of time. Convection and semiconvection are marked as indicated, and gray shading designates regions of nuclear energy generation. The upper solid line indicates the total mass of the star. The rapid phase of the Case~A mass transfer and the Case~AB mass transfer correspond to the sharp decreases of the total mass at about 4.9 and 5.9 Myr, respectively. Core hydrogen exhaustion coincides roughly with Case~AB mass transfer. Core helium burning ends at $t\simeq 7.1\,$Myr. The computations stops at core neon ignition. The decrease in mass between 4.9 and 5.9 Myr (about 2.5$\mso$) is due to slow Case~A mass transfer, that for $t\gtrsim 6\,$Myr (about 3.2$\mso$) due to WR winds. The final mass of the star before it explodes as supernova is 3.17$\mso$ (cf.~Table~1). } \end{centering} \label{fig:m26conv} \end{figure*} \begin{figure} \epsfxsize=\hsize \epsffile{8742f3.ps} \caption{Details of the mass transfer evolution of our progenitor model for Wray~977/GX~301-2, System No.~8. The left panel covers the Case~A mass transfer phase, the right panel the Case~AB. Shown are: the mass transfer rate (upper panel), stellar radius $R_{\rm 1}$ (solid line) and Roche radius $R_{\rm L,1}$(dashed line) of the primary (middle panel), and stellar radius $R_{\rm 2}$ (solid line) and Roche radius $R_{\rm L,2}$(dashed line) of the secondary (lower panel).}\label{fig:m_trans} \end{figure} \begin{figure}[t] \epsfxsize=\hsize \epsffile{8742f6.ps} \caption{Tracks in the HR diagram for the secondaries of the Systems No.~10 ($\alpha_{\mathrm{sc}}=0.01$, dashed-dotted line), No.~10a ($\alpha_{\mathrm{sc}}=0.02$, dashed line), and No.~10b ($\alpha_{\mathrm{sc}}=0.04$, solid line), which have the same initial stellar masses of 25 + 24 $\mso$ and the same initial period of 3.5~d (cf. Table~1). For Systems No.~10 and~10a, the tracks cover the evolution of the star from the zero age main sequence until the supernova explosion, reverse mass transfer does not occur. In System No.~10b, the secondary star rejuvenates during core hydrogen burning; reverse mass transfer would occur but is not taken into account in this calculation; the track ends before core helium ignition. }\label{fig:m25hrd} \end{figure} We emphasize that our progenitor model for Wray~977 /GX~301-2, in particular the initial mass of the neutron star progenitor as well as its final CO-core mass, depend only very weakly on the convection criterion or convective core overshooting. This is particularly true for the implication that the primary component forms a neutron star since, in contrast to single stars, the final masses of the mass loser in massive binaries (${\mathrm M > 25 \mso}$) are determined by the WR wind mass loss (see Sections~4.1 and~4.2, and Figure~\ref{fig:mco-mi} below). Finally, we want to mention that the surface composition of Wray~977 might be an additional way to discriminate evolutionary scenarios for its progenitor evolution. While according to scenarios which propose a completely non-conservative evolution ($\beta =0$) the surface chemical composition would be unaltered, our model predicts a enrichment factor for nitrogen of 5.5 and depletion factors of carbon and oxygen of 0.33, and 0.75, respectively. All isotopes of Li, Be and B are totally depleted. \subsection{Potential problems of the progenitor model} Our progenitor model for Wray~977/GX~301-2 has two potential problems which we want to discuss here for completeness, even though they may turn out not to be essential. The first concerns effects which the stellar wind of the secondary star might have on the accretion efficiency parameter $\beta$. In principle, one could imagine a situation where the secondary's wind drags part of the material which has left the primary and is on its way to the secondary with it to infinity. However, detailed models for such a situation are not available. We are optimistic that this effect is not too important in our case since most of the mass transfer from the primary to the secondary does not occur through an accretion disk, but instead, according to the estimates of Ulrich \& Burger (1976), the gas stream impacts directly onto the surface of the secondary. This is so for the complete Case~A mass transfer as well as for the ensuing Case~AB mass transfer in our System No.~8. Thus, the Case~A binaries are those among the massive close binaries for which a conservative evolution appears most appropriate. The second potential problem is that for certain initial and physical parameters the supernova order in Case~A systems can reverse, i.e., the secondary can explode before the primary does. We have shown in Section~3.2 that Case~A models exist for which the supernova order is not inverted (as needed to explain Wray~977/GX~301-2; e.g. our model No.~8). However, with $\alpha_{\rm sc}=0.01$ a reverse supernova order would most likely occur in this system, as we conclude from our model No.~10 (cf. Table~1). This does not mean that for $\alpha_{\rm sc}=0.01$ a supernova order reversal would occur for all conservative Case~A systems; rather a slight decrease of the initial secondary mass or a slight increase of the initial period would avoid this to happen. It is therefore conceivable that for any value of $\alpha_{\rm sc}$ the initial period-initial mass ratio parameter space for Case~A systems for which the primary explodes first as a supernova is not empty (cf. also Pols 1994; Wellstein et al. 1999). \subsection{Alternative progenitor scenarios for Wray~977} During our literature search, we did not find any other evolutionary calculations which were tailored to fit the system Wray~977/GX~301-2. Several ``scenarios'' for the progenitor evolution of this system have been considered. Brown et al. (1996), relying on the mass determination of $\gtrsim 48\mso$ for Wray~977 by Kaper et al. (1995), prefer 45$\mso$ as the progenitor mass of GX~301-2. They do not consider constraints imposed by the binary period. In Vanbeveren et al.'s (1998a) scenario the system consists initially of a $40 + 36\mso$ pair in a 50~d orbit, which, in a non-conservative ($\beta = 0.5$) Case~B mass transfer, evolves into an $18.5 + 41\mso$ helium star-O~star pair in a 28~d orbit. Interestingly, this is almost the same configuration which results from the detailed evolutionary calculations of de Loore \& De Greve (1992) starting also with a $40 + 36\mso$ pair but with a much smaller orbital period of 19.6~d. Finally, Ergma \& van den Heuvel (1998) prefer a completely non-conservative scenario ($\beta =0$) which implies an initial mass of GX~301-2 in excess of $50\mso$, from which they draw far reaching conclusions, in particular regarding the initial mass limit for black hole formation. In all of these scenarios, mass needs to be removed from the system. However, the processes which would do this are not well understood. Consequently, the amount of mass which leaves the system, and the amount of angular momentum which is removed together with the mass are not well constrained and need to be parameterized. This shows, e.g., when the above mentioned results of Vanbeveren et al. (1998a) and de Loore \& De Greve (1992) are compared, which differ in their treatments of angular momentum loss. Consequently, the period of Wray~977/GX~301-2, or likewise the orbital separation of the two components, can not be precisely predicted in these scenarios. Only the scenario of Ergma \& van den Heuvel (1998) does not include binary-induced mass loss from the system and is thus free of this uncertainty. All of the above scenarios favor an initial mass of the neutron star GX~301-2 in the range of 40...50$\mso$. Several of the papers mentioned above conclude that therefore single stars with initial masses in this range should form neutron stars as well. According to our detailed Case~A evolutionary model presented in Section~3.2 such conclusions can not be supported. On the contrary, we find a likely, and at least a possible, initial mass of GX~301-2 of only 26$\mso$. Furthermore, as will be outlined in Section~5 below, this implies only that single stars of initially 21$\mso$, not of 26$\mso$, form a neutron star. While this appears to be in comfortable agreement with (admittedly uncertain) expectations from single star calculations (e.g., Woosley \& Weaver 1995), neutron stars from 40...50$\mso$ stars are not, which made Ergma \& van den Heuvel (1998) conclude that the type of remnant of the mass loser --- neutron star or black hole --- must depend on additional stellar parameters, for example magnetic fields or rotation. Although we can not rule out that such effects play a role, we can not conclude from our results that such effects should be present. \begin{table*}[tp!] \caption{Key properties of the computed binary systems. Given are the initial primary and secondary masses $M_{\rm 1,i}$ and $M_{\rm 2,i}$, the initial mass ratio $q_{\rm i} = M_{\rm 2,i} / M_{\rm 1,i}$, initial and final period $P_{\rm i}$ and $P_{\rm f}$, the maximum mass of the secondary $M_{\rm 2,max}$, the mass transfer case, initial and final helium core mass of the primary $M_{\rm He,i}$ and $M_{\rm He,f}$ (note that the latter is equal to the final mass of the primary component for all considered cases), final CO-core mass of the primary $M_{\rm CO,f}$, the central carbon mass fraction of the primary at core helium exhaustion C$_{\rm c}$, the primaries final helium and carbon surface mass fraction He$_{\rm s}$ and C$_{\rm s}$, and the amount of helium left in the pre-supernova structure of the primary $\Delta M_{\rm He}$. Models 2$^{\prime}$, 5$^{\prime}$, 7$^{\prime}$, 10$^{\prime}$, 15$^{\prime}$ and 17$^{\prime}$ are computed with the WR-wind mass loss rates multiplied by a factor 0.5 (Eq.~2 in Section~2), that for System No.~1$^{\prime\prime}$ with 0.25. No final periods have been derived for these systems. All systems have been computed with a semiconvection parameter of $\alpha_{\rm sc} = 0.01$ unless indicated otherwise. System No.~8 corresponds to our progenitor model for the massive X-ray binary Wray~977/GX~301-2 and is discussed in detail in Section~3.2} \label{fig:systems} \begin{tabular}{l c c c c l l c c l l c c c l} \hline ~ & $M_{\rm 1,i}$ & $M_{\rm 2,i}$ & $q_{\rm i}$& $P_{\rm i}$ & $P_{\rm f}$ & $M_{\rm 2,max}$ & Case & $M_{\rm He,i}$ & $M_{\rm He,f}$ & $M_{\rm CO,f}$ & C$_{\rm c}$ & He$_{\rm s}$ & C$_{\rm s}$ & $\Delta M_{\rm He}$ \\ & $\mso$ & $\mso$ & & d & d & $\mso$ & & $\mso$ & $\mso$ & $\mso$ & & & & $\mso$ \\ \hline 1$^a$ & 60 & 34 &0.57& 20 & 59$^a$ & 58.9$^a$& B & 26.8 & 3.13 & 2.35 & 0.35 & 0.41 & 0.48 & 0.24 \\ 1$^{\prime\prime}$$^a$&60&34&0.57&20& --- & --- & B & 26.8 & 7.55 & 5.93 & 0.25 & 0.14 & 0.47 & 0.12 \\ 2 & 60 & 34 &0.57& 6.2 & 17.9 & 59.6 &A+AB & 25.8 & 3.10 & 2.34 & 0.35 & 0.50 & 0.42 & 0.29 \\ 2$^{\prime}$ & 60 & 34 &0.57& 6.2 & --- & 59.6 &A+AB & 25.8 & 4.07 & 3.07 & 0.32 & 0.28 & 0.53 & 0.17 \\ 3 & 60 & 40 &0.67& 7.0 & 18.9 & 64.7 & A+AB & 26.1 & 3.11 & 2.35 & 0.35 & 0.50 & 0.42 & 0.29 \\ 4 & 46 & 34 &0.74& 5.0 & 18.0 & 55.9 & A+AB & 18.2 & 3.10 & 2.34 & 0.35 & 0.48 & 0.44 & 0.27 \\ 5 & 40 & 30 &0.75& 4.0 & 16.5 & 51.2 & A+AB & 14.2 & 3.11 & 2.33 & 0.36 & 0.55 & 0.38 & 0.33 \\ 5$^{\prime}$ & 40 & 30 &0.75& 4.0 & --- & 51.2 & A+AB & 14.2 & 3.84 & 2.87 & 0.33 & 0.34 & 0.51 & 0.20 \\ 6$^a$ & 40 & 25 &0.63& 40 &130$^a$ &45.0$^a$& B & 16.9 & 3.11 & 2.34 & 0.35 & 0.50 & 0.42 & 0.29 \\ 7 & 30 & 24 &0.80& 4.0 & 22.0 & 41.9 & A+AB & 9.92 & 3.12 & 2.33 & 0.36 & 0.59 & 0.36 & 0.36 \\ 7$^{\prime}$ & 30 & 24 &0.80& 4.0 & --- & 41.9 & A+AB & 9.92 & 3.63 & 2.71 & 0.34 & 0.93 & 0.05 & 0.33 \\ {\bf 8}$^b$ &{\bf 26}&{\bf 25}&{\bf 0.96}&{\bf 3.5}&{\bf 46.2}&{\bf 42.4}&{\bf A+AB}&{\bf 6.41}&{\bf 3.17}&{\bf 2.33}&{\bf 0.37 }&{\bf0.98}&{\bf0.00}&{\bf 0.62} \\ 9 & 25 & 24 &0.96& 5.0 & 34.9 & 39.2 & B & 8.32 & 3.12 & 2.32 & 0.36 & 0.96 & 0.02 & 0.39 \\ 10$^c$ & 25 & 24 &0.96& 3.5 & 44.1 & 40.6 & A+AB & 6.02 & 3.19 & 2.36 & 0.37 & 0.98 & 0.00 & 0.66 \\ 10a$^b$ & 25 & 24 &0.96& 3.5 & 47.6 & 40.6 & A+AB & 6.06 & 3.19 & 2.31 & 0.37 & 0.98 & 0.00 & 0.71 \\ 10b$^b$ & 25 & 24 &0.96& 3.5 & --- & 40.6 & A+AB & 6.06 & --- & --- &--- & --- & --- & --- \\ 10$^{\prime}$ $^c$&25&24&0.96& 3.5 & --- & 40.6 & A+AB & 6.06 & 3.39 & 2.42 & 0.36 & 0.98 & 0.00 & 0.79 \\ 11 & 25 & 19 &0.76& 4.0 & 35.7 & 35.9 & A+AB & 6.27 & 3.18 & 2.32 & 0.37 & 0.98 & 0.00 & 0.71 \\ 12 & 25 & 16 &0.64& 4.0 & 28.3 & 32.3 & A+AB & 6.19 & 3.16 & 2.24 & 0.37 & 0.98 & 0.00 & 0.78 \\ 13 & 22 & 18 &0.82& 3.0 & 49.2 & 34.2 & A+AB & 4.48 & 3.13 & 1.90 & 0.37 & 0.98 & 0.00 & 1.12 \\ 14 & 21 & 19 &0.90& 5.0 & 43.4 & 33.2 & B & 6.00 & 3.14 & 2.03 & 0.36 & 0.98 & 0.00 & 0.98 \\ 15 & 21 & 19 &0.90& 3.0 & 60.2 & 34.4 & A+AB & 4.20 & 3.15 & 1.80 & 0.37 & 0.98 & 0.00 & 1.20 \\ 15$^{\prime}$ & 21 & 19 &0.90& 3.0 & --- & 34.4 & A+AB & 4.20 & 3.45 & 1.80 & 0.36 & 0.98 & 0.00 & 1.54 \\ 16$^b$ & 21 & 19 &0.90& 3.0 & 57.7 & 34.5 & A+AB & 4.23 & 3.07 & 1.73 & 0.35 & 0.98 & 0.00 & 1.25 \\ 17 & 20 & 18 &0.90& 5.0 & 45.8 & 31.7 & B & 5.56 & 3.15 & 2.17 & 0.37 & 0.98 & 0.00 & 0.73 \\ 17$^{\prime}$ & 20 & 18 &0.90& 5.0 & --- & 31.7 & B & 5.56 & 3.39 & 2.18 & 0.36 & 0.98 & 0.00 & 0.95 \\ 18 & 20 & 18 &0.90& 3.5 & 55.2 & 32.5 & A+AB & 4.30 & 3.00 & 1.69 & 0.37 & 0.98 & 0.00 & 1.23 \\ 19 & 20 & 16 &0.80& 3.5 & 47.8 & 30.7 & A+AB & 4.24 & 3.03 & 1.75 & 0.37 & 0.98 & 0.00 & 1.18 \\ 20 & 20 & 14 &0.70& 3.5 & 48.4 & 28.9 & A+AB+ABB& 4.20 & 2.71 & 1.58 & 0.37 & 0.98 & 0.00 & 1.06 \\ 21 & 20 & 18 &0.90& 2.5 & 65.5$^d$& 33.2 & A+AB+ABB& 3.55 & 2.7$^d$ & 1.56$^d$& 0.36 & 0.98 & 0.00 & 1.07$^d$ \\ 22 & 19 & 17 &0.90& 5.0 & 46.7 & 30.2 & B & 5.18 & 3.15 & 2.12 & 0.36 & 0.98 & 0.00 & 0.90 \\ 23 & 18 & 17 &0.94& 5.0 & 49.3 & 29.5 & B & 4.90 & 3.17 & 1.93 & 0.36 & 0.98 & 0.00 & 1.13 \\ 24 & 18 & 16 &0.89& 5.0 & 45.7 & 28.6 & B & 4.89 & 3.15 & 1.77 & 0.36 & 0.98 & 0.00 & 1.26 \\ 25 & 18 & 16 &0.89& 3.0 & 167 & 29.7 & A+AB+ABB& 3.27 & 2.02 & 1.48 & 0.38 & 0.98 & 0.00 & 0.51 \\ 26 & 16 & 15 &0.94& 8.0 & 172.1 & 25.7 & B+BB & 3.83 & 2.32 & 1.51 & 0.36 & 0.98 & 0.00 & 0.74 \\ 27 & 16 & 11 &0.69& 3.0 & 264$^d$& 24.2 & A+AB+ABB& 2.63 & 1.45$^d$ & 1.21$^d$& 0.40& 0.98 & 0.00 & 0.15$^d$ \\ 28 & 13 & 12 &0.92& 3.1 &175$^d$ & 22.0 & B+BB & 2.80 & 1.42$^d$ & 1.31$^d$& 0.40& 0.98 & 0.00 & 0.11$^d$ \\ \hline \end{tabular} $^a$ system was treated conservatively although a common envelope phase is expected. The final period and the maximum secondary mass may thus be greatly overestimated \\ $^b$ sequence was computed with $\alpha_{\rm sc}=0.02$, except for System No.~10b, where $\alpha_{\rm sc}=0.04$ was used \\ $^c$ reverse supernova order (no influence on core masses)\\ $^d$ masses are only upper ($M_{\rm He,f}$, $\Delta M_{\rm He}$) or lower ($M_{\rm CO,f}$) limits, and the final period is only a lower limit, since the Case~BB or~ABB mass trasfer phase was not finished at the end of the calculations. \\ \end{table*} \section{The final stellar and core masses of massive primaries } In this Section, we present the results of calculations for the evolution of massive close binaries for a wide range of primary star initial masses and for various initial system parameters. The evolution of the primary components is followed up to central neon ignition in most cases. Our goal is to derive final properties of the primaries, in particular their final masses and core masses. Those are used in the next Section to constrain the critical initial mass limit for neutron star/black hole formation in {\em single stars}. As a by-product we also obtain results for the evolution of the surface chemical composition of the primaries, which is relevant for the understanding of Type~Ib/c supernovae; these issues are discussed in Section~6. We compute binary models for primary star initial masses from 13 to 60$\mso$, and for various initial orbital periods and mass ratios. We also study the differences between Case~A and Case~B mass transfer for the resulting core masses. Since there are indications that the mass loss rates for Wolf-Rayet stars (i.e., helium stars) have been overestimated in the past (Hamann \& Koesterke 1998), we computed sequences where our standard helium star mass loss rate (cf. Section 2) has been multiplied by 0.5, in one case even by 0.25. An overview over the grid of computed systems can be obtained from Table~\ref{fig:systems}. Our calculations are restricted to Case~A and early Case~B systems. The latter are defined as such where the mass transfer occurs early enough during the post-main sequence expansion of the primary that contact is avoided. The theory for the evolution of late Case~B and Case~C systems is not yet very well developed; often it is assumed that such system go through a common envelope evolution and either merge --- in this case they are not relevant for the conclusions of our work --- or expell the hydrogen-rich envelope of the primary during the spiral-in phase of the secondary (e.g., Podsiadlowski 1992). Since this process also removes considerable amounts of orbital angular momentum rather small orbital separations and periods are expected from this type of evolution. Three types of post-main sequence evolutionary tracks can be distinguished. The first (say, Type~1) leads the star to the red supergiant branch at the begining of core helium burning but allows the stellar radius to increase after core helium exhaustion to values which exceed those achieved earlier. Tracks like this are in fact common, but the post-helium burning radius excess is mostly not very large, in particular for stars with initial masses above 20$\mso$ (e.g., Schaller et al. 1992). The second type of track (Type~2) reaches the Hayashi-line only at the end of core helium burning. However, this kind of evolution is found only for small metallicity (cf., Schaller et al. 1992, Langer \& Maeder 1994), and the observed large number of red supergiants discards it as a common case. There is also an intermediate type of evolutionary track (Type~3), where the star reaches the Hayashi line for the first time {\em during} central helium burning (e.g., the 20$\mso$ track at Z=0.02 or the 60$\mso$ track at Z=0.001 of Schaller et al. 1992). Unfortunately, it is not possible at the present time to correctly predict stars of which mass and metallicity follow which type of track. The reason is that the temperature and radius evolution of stars more massive than $\sim 10\mso$ is an extremely sensitive function of internal mixing efficiencies (cf. Stothers \& Chin 1992, Langer \& Maeder 1994) or mass loss rates (cf. Schaller et al 1992 and Meynet et al. 1994), which suffer from large uncertainties. Note that this problem must lead to large uncertainties in any prediction of the number of black hole binaries, which is not always adequately emphasized in population synthesis studies. Since in the present paper we deal only with stars of roughly solar metallicity, we neglect the possibility of Type~2 tracks in the following discussion. We also negect the unusual Type~3 tracks; however, as we can not prove that they are not common, we note here that the initial-final mass relation for primaries of this type would be inbetween that found for the Case~A/B systems and that of single stars (cf. Fig.~5 below). With these assumptions, the evolution of the primaries for systems which do not merge or experience reverse mass transfer depends only on the time during their evolution when the hydrogen-rich envelope is removed. Consequently, the primaries of late Case~B systems evolve like primaries of early Case~B systems of the same initial mass. Furthermore, as in Case~C systems the mass transfer starts only after core helium exhaustion, and stellar wind mass loss after the mass transfer can be neglected due to the short remaining stellar life time, the helium cores of the primaries in these systems evolve like the helium cores of single stars of the corresponding initial mass. \subsection{Initial-final mass relations for primaries} The results obtained for the initial-final mass relations of the primaries in massive close binaries are shown in Figure~\ref{fig:mco-mi}a. The data for Case~A and Case~B systems are taken from Table~1, where the primaries final helium core mass is designated as $M_{\rm He,i}$. Since all of them are devoid of hydrogen in their pre-supernova stage, the final helium core mass is equal to their final mass. Except for Systems No.~21,~27 and~28, mass loss occuring beyond the end of our calculations (neon ignition in most cases) can safely be neglected due to the short remaining life time of the star. As expected from earlier work, the initial and therefore also the final helium core masses of Case~B primaries do basically not depend on the initial mass ratio or the initial orbital period (e.g., de Loore \& De Greve 1992). Simply, the mass transfer stops only when almost all of the hydrogen-rich envelope is removed from the primary. However, since the period evolution does depend on the two mentioned initial parameters, this simple logic does not apply any more to primaries which evolve to final masses below $2.5...3\mso$, since helium stars with masses smaller than this tend to evolve into red giants (cf., Habets 1986) which then may or may not lead to a so called Case~BB mass transfer (Delgado \& Thomas 1981). We do not investigate this process in detail here since it turns out that it does not affect the initial mass limit for neutron star/black hole formation. It is certainly important, though, for investigations of the initial mass limit for white dwarf/neutron star formation in binary systems. Figure~\ref{fig:mco-mi}a shows that the initial-final mass relation of Case~A systems differs from that of Case~B systems for initial primary masses below $\sim 20\mso$. The reason is that Case~A mass transfer leads to smaller initial helium core masses compared to Case~B, which has been demonstrated in detail in Section~3.2 at the example of our progenitor model for Wray~977. Note that there is no strict initial-final mass relation for Case~A systems at all since the initial helium core mass depends on time during core hydrogen burning when the mass transfer starts. Smaller initial helium star masses are thus obtained for smaller initial periods (compare, e.g., systems No.~18 and ~21). A dependence on the initial mass ratio seems to be very small or absent (cf. Systems No.~18, 19, and~20). That the initial-final mass relations for Case~A and~B primaries are almost identical above $\gtrsim 25\mso$ is due to the fact that, for our standard mass loss rate (Equation~1 in Section~2), the final masses become independent of the initial helium star mass (Langer 1989). Even when the Wolf-Rayet mass loss rate is reduced by a factor of~2 (Equation~2 in Section~2), the initial-final mass relation remains rather flat, although a slight positive slope is found in this case (Fig.~\ref{fig:mco-mi}a). Due to this mass convergence, also the efficiency of convective core overshooting on the initial-final mass relation above $\gtrsim 25\mso$ can be assumed to be small. Fig.~\ref{fig:mco-mi}a shows also the initial mass-final helium core mass relations for single stars obtained from the literature which, as mentioned above, might also aply to Case~C mass transfer systems. It is surprising that the treatment of convection, in particular the so called convective core overshooting, introduces a much larger uncertainty to the initial-final mass relation of single stars than to that of Case~A/B binaries. The models of Langer \& Henkel (1995), which have been computed with the same treatment of convection as the models presented here, predict smaller final helium core masses for stars below $\sim 30\mso$ but larger ones in the mass range 30...60$\mso$. For much higher initial masses, both sets of models can be assumed to converge, as the curve derived from the models of Maeder (1992) and that of the Case~A/B binaries do, due to the effect of mass convergence (all computations use very similar Wolf-Rayet mass loss rates). \begin{figure*}[p] \begin{centering} \vbox{ \epsfxsize=0.95\hsize \epsffile{8742f4.ps} } \end{centering} \caption{{\bf a)} Final He-core masses of the primaries of massive close binaries as function of their initial mass, for Case~A and Case~B systems (cf. Table \ref{fig:systems}) for nominal Wolf-Rayet mass loss (solid lines, below 25 $\mso$ the upper line marks Case~B systems) and --- for the Case~A --- also for half the nominal Wolf-Rayet mass loss rate (dotted line), compared to the final helium core masses of single stars (dashed line; cf. Langer \& Henkel 1995; dotted-dashed line; cf. Maeder 1992) which may give a good approximation to the non-merging fraction of the Case~C systems. {\bf b)}: The same as Fig.~\ref{fig:mco-mi}a, but for the final CO-core masses. The gray horizontal line marks the final CO-core mass of our progenitor model for the neutron star companion to Wray~977, GX~301-2} \label{fig:mco-mi} \end{figure*} \subsection{Initial mass - final CO-core mass relations} In order to draw conclusions for the initial mass limit for the formation of neutron stars/black holes, the initial-final mass relation is not sufficient, since --- due to the strong mass loss, particularly in binary systems --- stars with the same final mass or even with the same final helium core mass may evolve different final CO-core masses. This can be seen in Figure~\ref{fig:mco-mi}b, where we plot the CO-core masses as function of the initial mass for the same models as in Fig.~\ref{fig:mco-mi}a. E.g., the Case~A and~B primaries with initial masses between 20~and 25$\mso$ end up with the same final helium core mass but with largely different CO-core masses. Above initial primary masses of $\sim 25\mso$, the final CO-core masses are determined by the Wolf-Rayet winds. As in the case of the final helium core masses, the slope of the initial mass-final CO-core mass relation is zero for the larger Wolf-Rayet mass loss rate and slightly positive for the smaller mass loss rate. Again, the mass transfer type (A~or~B) or the treatment of convection can be assumed to have very little influence. \subsection{Constraints from massive X-ray binaries} Our 26$\mso$ progenitor for GX~301-2 (cf. Section~3.2) has a final CO-core mass of 2.33$\mso$. Note that the initial mass-final CO-core mass for Case~A/B primaries is particularly certain around this mass. While uncertainties in the treatment of convection become relevant only below $\sim 25\mso$ (cf. Section~4.2), significant differences in the final CO-core masses due to differences in the Wolf-Rayet mass loss rate occur only for higher initial primary masses. As 26$\mso$ is the lowest possible initial mass for GX~301-2 we can conclude that from Fig.~\ref{fig:mco-mi} that Case~A and~B primaries with initial masses equal or less than this do not form black holes, but instead evolve into neutron stars or white dwarfs. On the other hand, the flatness of the initial mass-final CO-core mass relation for Case~A/B primaries makes it very hard to form black holes for {\em any} initial primary mass, and impossible to form high mass black holes (cf. Section~5.2). In fact, for the larger Wolf-Rayet mass loss rate any black hole formation would be excluded. For the reduced Wolf-Rayet mass loss rate it may be possible to form black holes, but it would be low mass black holes according to our definition in Section~5.2 (cf. Fig.~\ref{fig:mco-mi}a). We computed one massive system with a 60$\mso$ primary aplying only 1/4th of the Wolf-Rayet mass loss rate (model No.~1$^{\prime\prime}$; cf. Table 1). The primary ended as a 7.5$\mso$ Wolf-Rayet star. A 36$\mso$ helium core --- which may correspond to a 85$\mso$ star --- computed with the same mass loss rate ended with 7.9$\mso$. We conclude that, with the present stellar wind mass loss rates, it appears to be impossible to explain a system like Cygnus~X-1 with a black hole mass of $\sim 10\mso$ (Gies \& Bolten 1982, 1986; Herrero et al. 1995) through Case~A or Case~B mass transfer. Also the other two potential massive black hole binaries, LMC~X-3 with a black hole mass of $5.5\pm 1\mso$ (Kuiper et al. 1988), and LMC~X-1 with a probable black hole mass of 6$\mso$ (Hutchings et al. 1987) are not likely to be Case~A or~B remnants. Only if the Wolf-Rayet mass loss rates turn out to be substantially smaller than one fourth of Eq.~1, a Case~A/B evolution for the progenitor of Cygnus~X-1 will be possible, as suggested by Vanbeveren et al. (1998bc). Anyway, a Case~C evolution appears possible for this system. E.g., for Cygnus~X-1, Gies \& Bolten (1986) find that the mass of the optical component should be $\gtrsim 20\mso$, while Herrero et al. (1995),by performing a detailed spectral synthesis, find a most likely mass of 17.8$\mso$. Thus, the progenitor of the black hole could have had an initial mass $\gtrsim 25\mso$, which is consistent with a large black hole mass according to Fig.~5a. Interestingly, Brown et al. (1999) have recently proposed Case~C mass transfer (mass transfer after central helium exhaustion) as a possibility to obtain massive ($\sim 7\mso$; cf. Baylin et al. 1998) black holes in the six known low mass black hole binaries; in fact, due to the extrapolated large population of Galactic low mass black hole binaries (Romani 1998), Portegies~Zwart et al. (1997) found alternative explanations extremely difficult. According to Fig.~\ref{fig:mco-mi}a, these black holes should then come from an intermediate mass range, i.e. roughly 25...50 $\mso$, since the most massive Case~C primaries, like single stars, are supposed to evolve through a Luminous Blue Variable phase (Langer et al. 1994) where they lose the major part of their hydrogen-rich envelope before core helium ignition and quickly transform into Wolf-Rayet stars. They therefore would most likely avoid mass transfer altogether (Vanbeveren 1991, Langer 1995). According to Fig.~\ref{fig:mco-mi}a, the maximum black hole initial mass in Case~C systems, assuming that the hydrogen-rich envelope is completely expelled at the end of the spiral-in process, is of the order of 10$\mso$. Lower stellar wind mass loss rates could lead to even higher initial black hole masses. From Figure~\ref{fig:mco-mi} it seems also conceivable that the average black hole mass in black hole binaries is rather large, i.e., well above the final masses of Case~A/B primaries, as suggested by Baylin et al. (1998). A distinction of two different regimes of remnant masses has already been elaborated by Brown et al. (1996). While it can not be excluded that the most massive Case~A/B primaries form black holes of relatively low mass ($\sim 3\mso$), their number is not expected to be very large and is possibly zero (Figure~\ref{fig:mco-mi}b). On the other hand, assuming that in a failed supernova the whole helium core forms the initial black hole (cf. Fryer 1999) implies that most of those which form in Case~C systems will have initial masses in excess of $\sim 5\mso$ (Figure~5a). \section{Transforming the mass limit from binaries to single stars} \subsection{Assumptions} For single stars which do not lose their hydrogen-rich envelope completely it is well known that the evolution of the stellar core is decoupled from the envelope evolution. This allows the study of the inner properties of stars during their late evolutionary phases --- and thus also the investigation of their fate --- by just computing the evolution of helium cores of a given mass (Arnett 1977, Woosley \& Weaver 1988, Thielemann et al. 1996). For stars which do lose their hydrogen-rich envelope well before core helium exhaustion, i.e., in particular the primaries of massive close binary systems, this approximation is not possible any more because the mass of the helium core decreases during core helium burning. However, as the mass loss never reaches the next inner core, the CO-core, we will make the assumption in the following that the final CO-core mass determines the fate of a massive star. In particular we assume that CO cores with a mass below a critical value form neutron stars or white dwarfs, while more massive CO-cores form black holes. We note that the central carbon abundance at core helium exhaustion $C_{\rm c}$ is in principle a futher independant parameter to be considered. It determines whether core carbon burning occurs radiatively or convective, which affects the core entropy and the final iron core mass (cf. Woosley \& Weaver 1995, Brown et al. 1999). As more massive stars perform core helium burning at higher temperatures, the general trend is that $C_{\rm c}$ is smaller for higher initial masses. There appears to be a critical carbon mass fraction of about 15\% below which carbon burning is radiative (Weaver \& Woosley 1993, Woosley \& Weaver 1995, Heger et al. 1999), and the initial mass-final iron core mass-relation jumps discontinously to a larger value (Timmes et al. 1996 Brown et al. 1999). The situation is complicated by the fact that $C_{\rm c}$ is not only dependant on the initial stellar mass. For stars which lose the hydrogen-rich envelope during their evolution and uncover their helium core --- i.e. in particular the primaries of massive close binaries ---, a higher mass loss during core helium burning leads to larger values of $C_{\rm c}$ (Woosley et al. 1995, cf. also Table~1). Even though Woosley \& Weaver (1995) find stars more massive than $\sim 19\mso$ fall short of the critical carbon value of $\sim 15\%$, recent models of Heger et al. (1999) which include effects of rotational mixing find higher values of $C_{\rm c}$ for single stars up to 25$\mso$. Langer (1991) has shown that, for given initial mass and mass loss history, $C_{\rm c}$ depends sensitively on assumptions about convection (cf. Table~1 in Langer 1991). Furthermore, it depends on the still poorly known $^{12}$C($\alpha ,\gamma$) nuclear reaction rate (Hoffman et al. 1999). Consequently, the values of $C_{\rm c}$ must be considered as uncertain. However, the central carbon mass fractions of our primaries are all well above the critical value of $\sim 15\%$ (cf. Table~1). We can thus assume a smooth CO-core mass-final iron core mass-relation for our models (neglecting the statistical fluctuations in such a relation due to silicon shell burning episodes; cf. Woosley \& Weaver 1995). Furthermore, as all our models have extended convective core carbon burning phases. As outlined by Brown et al. (1999) this results in rather low core entropies and and relatively small iron core masses. All together, we have reasons to assume that $C_{\rm c}$ is not an essential independent parameter affecting the results of the present study. In principle, also further parameters might be important, i.e, the stellar angular momentum or magnetic field; those are not considered here. \subsection{Results} The approximation often made for single stars that the fate depends monotonously on the initial stellar mass is not generally applicable for close binary components. This can be seen in Figure~\ref{fig:mco-mi}b, which shows that the final CO-core mass (and thus the fate) for a given initial mass depends on the previous mass transfer evolution. The progenitor of the neutron star in Wray 977/GX~301-2 must have had a CO-core mass of at least 2.33$\mso$, corresponding to the minimum initial mass of the primary of 26$\mso$. Fig.~\ref{fig:mco-mi}b shows that single stars above 21$\mso$ form CO-cores of at least 2.33$\mso$. This allows us to derive a minimum black hole progenitor mass for single stars of 21$\mso$ from Fig.~\ref{fig:mco-mi}b. This initial mass limit refers to models computed with the Ledoux criterion for convection (Langer \& Henkel 1995). For single stars computed with the Schwarzschild criterion and overshooting (Maeder 1992) we derive a limit of 13 $\mso$ from the same arguments. Since the two assumptions about convection can be considered as the two extreme cases, we conclude that single stars with initial masses below 13...21$\mso$ form neutron stars. We want to emphasize that this is the strongest statement that, at the present time, can be derived from the existence of a neutron star companion to Wray~977. In particular, even though especially the lower limit of 13$\mso$ appears disappointingly un-constraining, we must consider arguments in the literature which derive much higher initial mass limits for neutron star formation from the system Wray~977/GX~301-2 as wrong. If the hitherto poorly investigated Case~C mass transfer in massive stars would always lead to the merging of both components, the presence of black holes in massive binaries, e.g. that in Cyg~X-1, would exclude the standard WR wind mass loss rates (dotted lines in Fig.~\ref{fig:mco-mi}), because using this we find no systems with more massive final CO-cores than for the neutron star progenitor of GX~301-2 (cf. also Ergma \&\ van den Heuvel 1998). For the reduced WR wind mass loss rate --- assuming that at least some of the mass losers in binaries reach massive enough CO-cores to collapse to a black hole --- we derive an upper CO-core mass limit up to which neutron star formation might be possible of 3$\mso$. This would correspond to a single star initial mass limit of 26$\mso$ with the Ledoux criterion and of 15$\mso$ with the Schwarzschild criterion and overshooting. These are lower mass limits for black hole formation in single stars {\em if}~Case~A/B binaries could be shown to produce any black hole at all. It is interesting to distinguish the delayed black hole formation due to fall back (e.g., Woosley \& Weaver 1995), thermal effects (Wilson et al. 1986, Woosley \& Weaver 1986), or effects of the equation of state (Brown \& Bethe 1994), and the prompt formation of black holes. While delayed black hole formation gives rise to a ``normal'' supernova explosion, prompt black hole formation does not, although also in this case the hydrogen-rich envelope may be ejected (MacFadyen \& Woosley 1999, Fryer 1999). In the first case, the black hole mass equals the mass of the initially formed neutron star plus the mass of the matter which falls onto it later-on. Thus, black holes of rather low mass may form like this (Brown \& Bethe 1994, Brown et al. 1996). For the prompt stellar collapse to a black hole, it may be a reasonable assumption that the black hole mass roughly equals the final helium core mass (MacFadyen \& Woosley 1999, Fryer 1999). Even though these assumptions are uncertain and their confirmation must await a better understanding of the core collapse supernova mechanism, it probably makes sense that promptly formed black holes have larger masses than those formed through the delayed scenarios (Brown \& Bethe 1994, cf., however, Woosley \& Weaver 1995). In the line of Brown \& Bethe (1994), we will thus consider the first as high mass black holes (HMBHs) with the mass of the progenitor's helium core mass, and the latter as low mass black holes (LMBHs) with a mass of $\lesssim 3\mso$. The observation that Supernova~1987A, which had a progenitor of $\sim 20\mso$, did not promptly form a black hole (Arnett et al. 1989) can give important constraints on the black hole formation limits. Considering the single star models computed with the Schwarzschild criterion and overshooting, we conclude from Fig.~5b that prompt black hole formation can only occur in CO-cores more massive than $\sim 5\mso$. This limits the formation of high mass black holes in single stars or Case~C binaries to the initial mass range 20...32$\mso$. Masses of high mass black holes of the range 5...10$\mso$ could be produced (Fig.~5a). Case~A/B binaries could not form high mass black holes, and even the production of low mass black holes might be excluded as the core mass range which is able to yield to delayed black hole formation is rather narrow (cf. Brown et al. 1999). If the existence of a neutron star in the remnant of SN~1987A were confirmed (cf., Wu et al. 1998, Fryer et al. 1999) the models using the Schwarzschild criterion and overshooting might have a problem, since only a very small initial mass range might be left to form high mass black holes. For models using the Ledoux criterion for convection, these mass limits work out quite differently. From the neutron star companion to Wray~977, which resulted in a single star initial mass limit for black hole formation of $> 21\mso$ (see above), it follows that SN~1987A formed a neutron star, in accord with current models and observations (Wu et al. 1998, Fryer et al. 1999). No further firm initial mass limit can be derived from Fig.~5b, neither for low nor for high mass black hole formation. However, as stars above 21$\mso$ form much more massive CO-cores than stars around 20$\mso$, in particular stars with initial masses above $\sim 30\mso$ (Fig.~5b), it is conceivable that single stars and Case~C primaries above $\sim 30\mso$ or so form high mass black holes. As for the models computed with the Schwarzschild criterion, Case~A/B binaries would not produce high mass black holes. For the case that {\em all} black holes in close binaries form through the Case~C channel, we can only derive a less stringent lower mass limit for black hole formation as for the case that {\em some} black holes from in Case~A/B binaries. The limit then comes from the fact that the Case~C, at solar metallicity, is restricted to initial primary masses of less than $\sim 40\mso$ (cf. Sect.~4.3). The large number of low mass black hole binaries (Portegies Zwart et al. 1997, Romani 1998) with massive black holes ($\sim 7\mso$; cf. Baylin et al. 1998) makes it plausible that single stars in the mass range 25...40 $\mso$ form high mass black holes. Thus, independent of whether black hole binaries form preferentially through Case~A/B or Case~C, we conclude that very likely single stars with initial masses above $\sim 25\mso$ form black holes. This value is in agreement with the recent result from core collapse simulations of Fryer (1999). \section{The structure of supernova progenitors} The massive close binary models discussed in the previous Sections do not only give predictions for the final stellar masses of both components, but they also yield the final mechanical and chemical structure of their envelopes. While those primaries of our systems which do not collapse directly to black holes become Type~Ib or~Ic supernovae, the final explosions of the secondaries are hydrogen-rich and have thus to be classified as Type~II supernovae (cf. Langer \& Woosley 1996). The mechanical structure of the secondaries of massive close binaries can differ appreciably form single star Type~II supernova progenitors. While the latter are red supergiants, our secondaries are typically blue supergiants in their final stages. Whether or not massive secondaries explode as blue or red supergiants depends critically on the semiconvective mixing efficiency during the accretion phase (cf. Fig.~4), i.e., whether or not the secondary rejuvenates (Hellings 1983, Braun \& Langer 1995). In fact, for the the semiconvective mixing efficiency adopted here we can estimate a number ratio of Type~II supernovae from blue supergiants to Type~Ib/c supernovae of order unity. This does not contradict the fact that many more Type~Ib/c supernovae are observed, since exploding blue supergiants are about 4~magnitudes dimmer than Type~Ib supernovae (Young \& Branch 1989), as demonstrated by Supernova~1987A (Arnett et al. 1989). As mentioned above, the primaries of the systems studied in this work will mostly produce Type~Ib/c supernovae. As a main criterion to distinguish Type~Ib and Type~Ic supernovae is the detection of helium lines in Ibs but their absence in Ics (Harkness \& Wheeler 1990), we list in Table~\ref{fig:systems} the remaining mass of helium in the envelope of the primaries at the time of their supernova explosion. Figure~\ref{fig:dmhe} shows this quantity plotted as a function of the primaries initial mass, separately for Case~A and Case~B systems. It shows that substantial amounts of helium (i.e. more than $0.5\mso$) remain only in a limited initial mass range, i.e. above $\sim 15\mso$ and below $\sim 25\mso$. Progenitors with lower or higher initial masses, although they do not become helium free, may end up with as little as $\sim 0.1\mso$ of helium. For progenitors below $\sim 15\mso$, the reason for the low amounts of helium in the pre-supernova structure is the occurrence of a Case~BB or Case~ABB mass transfer as the low mass primaries, at the end of their helium-star phase, extend to red giant dimensions (cf. Section~4.1). These objects end with a total mass very close to the Chandrasekhar mass. For the Case~A, this happens in our models for primary initial masses below 18 $\mso$, although even the 20 $\mso$ primaries undergo a brief Case~ABB mass transfer which is, however, too short to remove significant amounts of mass. A comparable situation occurs in System No.~26, a Case~B system with a $16\mso$ primary: the Case~BB mass transfer occurs but stops before the major part of the helium envelope is removed. As it may have observational implications, we note that some of these objects explode {\em during} the mass transfer. The Type~Ib/c progenitors with initial masses above $\sim 25\mso$ are helium-poor due to strong stellar winds during a Wolf-Rayet phase. Figure~\ref{fig:dmhe} shows that this result is not strongly affected by uncertainties in the assumed Wolf-Rayet mass loss rate. {\em Smaller} amounts of helium remain for the most massive progenitors when the mass loss rate is reduced by a factor of~2. The reason is that the lower Wolf-Rayet mass loss leads to larger convective cores in the advanced helium burning stages. The final masses of these stars is well above the Chandrasekhar mass. Although the remaining amount of helium is similar for initial primary masses below $\sim 15\mso$ and above $\sim 25\mso$, the envelope mass, i.e. the amount of mass above the CO-core, is much larger in the latter. While the stars from below $\sim 15\mso$ have a mantle of pure helium (plus 2\% metals) on top of their CO-core, the envelope of the primaries from above $\sim 25\mso$ is strongly enriched with carbon and oxygen, as can be seen from the final surface abundances shown in Table~1. We could now speculate that the Type~Ic supernovae correspond to Case~A/B primaries with initial masses below $\sim 15\mso$ and above $\sim 25\mso$. As single stars above $\sim 40\mso$ evolve into Wolf-Rayet stars with correspondingly low remaining amounts of helium in the pre-supernova stage, those, together with Case~C primaries above $\sim 40\mso$, might as well contribute to the Type~Ic supernova class. In fact, small amounts of helium might still be compatible with Type~Ic supernovae (Filippenko et al. 1995). On the other hand, Case~A/B primaries from the initial mass range 15$\mso$...25$\mso$ as well as Case~C primaries from initial masses below $\sim 40\mso$ might evolve into Type~Ib supernovae. On the other hand, Woosley \& Eastman (1995) argued that the amount of helium seen in a Type~Ib/c supernova (i.e., the helium line strengths) may not only depend on the amount of helium which is present. They conclude that it is essential whether radioactive $^{56}$Ni is mixed close to or into the helium layer during the explosion, as the nickel decay can excite the helium line. Hachisu et al. (1991) have shown that more such mixing is expected in the explosions of lower mass helium stars. In summary, close binary primaries with the largest initial masses ($\gtrsim 25\mso$ for the Case~A/B) as well as single stars and wide binaries (denoted ``Case~C'', although no mass transfer occurs; cf. Section~4.3) with initial masses above $40...50\mso$ are the best candidates for Type~Ic supernova progenitors. In their pre-supernova stage, they contain low amounts of helium in their envelopes {\em and} they might not experience significant mixing of radioactive $^{56}$Ni into the helium layer. This is interesting in context with the peculiar Type~Ic supernova~1998bw and its associated weak $\gamma$-ray burst (Kulkarni et al. 1998). In the collapsar model of MacFadyen \& Woosley (1999), $\gamma$-ray bursts may be formed in helium cores which evolve a sufficiently massive iron core to undergo a prompt collapse to a black hole, given that it has the right amount of angular momentum. Independent of the convection physics, a prompt black hole formation {\em and} a Type~Ic supernova with small amounts of helium appear possible for Case~C binaries and also for single stars with initial masses of $\sim 40\mso$ or above. It is thus consistent with our conclusions that SN~1998bw is a borderline case of a successful $\gamma$-ray burst produced according to the model of MacFadyen \& Woosley (1999). \begin{figure}[t] \epsfxsize=\hsize \epsffile{8742f5.ps} \caption{The amount of helium present in the primary components at their immediate pre-supernova stage as function of their initial mass, for the computed Case~A (solid line) and Case~B (dotted line) systems. The dashed line corresponds to Case~A systems computed with a WR wind mass loss rate reduced by 50~\%.} \label{fig:dmhe} \end{figure} \section{The overall picture} In Table~2, we summarize a possible overall picture for the outcome of massive close binary evolution through the various mass transfer cases, and for single star evolution. One can, however, {\em not} pick initial primary mass and period (i.e., mass transfer case) and then obtain the final answer at the corresponding position in the table. Due to uncertainties in the convection efficiency, the Wolf-Rayet mass loss rate, the post-main sequence radius evolution of massive stars, and the CO-core-compact remnant mass relation, the dividing lines between the various entries in Table~2 can not yet be completely determined. Table~2 should rather be regarded as a diagram with the initial binary period as the x-axis and the initial primary or stellar mass as y-axis. Several {\em qualitative} trends can be found in this ``diagram'', which should be taken more seriously than the exact location of the dividing lines. \begin{table}[t] \caption{Type of remnant of primary components of massive close binaries, i.e., white dwarf (WD), neutron star (NS), low (LMBH) or high mass black hole (HMBH), and expected supernova type (if an explosion occurs, which is not secure for HMBHs) as function of the primary initial mass and the mass transfer type for a possible scenario of massive close binary evolution. The initial mass ranges are not to be taken too literally (see text). } \label{fig:cc} \begin{tabular}{ l l l l l } \hline \noalign{\vskip 3pt} M$_{1, \rm i}$ & Case A & Case B & Case C & single star \\ $\mso$ & & & & \\ \noalign{\vskip 3pt} \hline \noalign{\vskip 5pt} 8...13 & WD & WD & SN~Ib & SN~II \\ & & & NS & NS \\ \noalign{\vskip 15pt} 13...16 & WD & SN~Ib or~Ic & SN~Ib & SN~II \\ & & NS & NS & NS \\ \noalign{\vskip 15pt} 16...25 & SN~Ib & SN~Ib & SN~Ib & SN~II \\ & NS & NS & NS & NS \\ \noalign{\vskip 15pt} 25...40 & SN~Ic & SN~Ic & SN~Ib & SN~II \\ & NS & NS & HMBH$^a$& HMBH$^a$\\ \noalign{\vskip 15pt} $> 40$ & SN~Ic & SN~Ic & SN~Ic & SN~Ic \\ & NS/LMBH & NS/LMBH & HMBH$^b$ & HMBH$^b$ \\ & & &NS/LMBH$^c$ &NS/LMBH$^c$ \\ \noalign{\vskip 3pt} \hline \end{tabular} $^a$ mass range depends on assumptions on convection; see text\\ $^b$ for models computed with the Ledoux criterion for convection\\ $^c$ for models computed with the Schwarzschild criterion and convective core overshooting\\ \end{table} Most important, we see that the dividing lines between equal final stages are not horizontal but sloped, with a positive slope (for larger initial masses having smaller y-values) for all of them. For example, the dividing line between white dwarf and neutron star formation is (roughly) at 16...13$\mso$ for Case~A binaries (cf. System No.~27 in Table~1) --- it depends smoothly on the initial period for Case~A as shown by Wellstein et al. (1999) ---, at $\sim 13\mso$ for the Case~B (cf. System No.~28 in Table~1), and somewhere between~8 and~10$\mso$ for single stars and Case~C binaries. Similarly, high mass black holes are preferentially produced for large initial masses {\em and} large initial periods. Similar trends can be found for the two dividing lines between Type~Ib and~Ic supernovae (where the one at smaller initial masses may or may not correspond to the Type~Ib/Ic distinction; cf.~Section~6). This slope of the dividing lines between equal final stages are not small. For example, our models are consistent with an initial primary mass limit for black hole formation in Case~A/B systems of well above $100\mso$, while in Case~C systems and single stars we derive a value of $\sim 25\mso$. We thus emphasize that this effect should be included in population synthesis models which predict X-ray binary or $\gamma$-ray burst frequencies. For the same reason, the neutron star or black hole mass functions derived from single star models can not simply be compared with mass functions of compact companions in X-ray binaries. As discussed in Section~5.2, the predicted initial mass range for high mass black hole formation in Case~C binaries and single stars is quite different for models using the Ledoux criterion for convection ($\gtrsim 30\mso$) compared to models using the Schwarzschild criterion and overshooting (20...32$\mso$). In the latter case, the most massive single stars would evolve into low mass black holes or even neutron stars. This conclusion can be changed only if the Wolf-Rayet mass loss rate were smaller than one fourth of that in Eq.~1 (cf. Vanbeveren et al. 1998b). Finally, we note that the orbital periods of the Case~A,B,C binaries at the time when the primary has terminated its evolution occur in reverse order as the initial orbital periods (cf., Verbunt 1993). Early Case~B systems end with shorter periods than the Case~A systems because, for a given primary initial mass, less mass is transfered during a Case~B mass transfer compared to the amount of mass which is transfered in a Case~A and~AB together. Furthermore, while the conservative Case~A and~B binaries evolve to periods of several weeks and months --- our Wray~977/GX~301-2 progenitor model No.~8 with a final period of 46~days is a typical case --- late Case~B and Case~C systems lose substantial amounts of mass and angular momentum and thus become short period binaries (if they do not merge). E.g., the system 4U~1700-37, a $\sim 30\mso$ O~star with a 2.6$^{+2.3}_{-1.4}\mso$ compact companion and a 3.4~day orbital period (Heap \& Corcoran 1992, Rubin et al. 1996), must have gone through a non-conservative mass transfer. This implies that the progenitor mass of the compact object is likely to be larger than that of GX~301-2. Consequently, conservative Case~A and~B systems, which are modeled in detail in the present work, are the best candidates for progenitors of close binaries consisting either of a helium star and a compact remnant or of two compact stellar remnants --- i.e. for binary $\gamma$-ray burst progenitors (cf. Fryer et al. 1999, and references therein): Due to the large periods at the end of the primary's evolution the binding energy of the secondary's envelope is low when the reverse mass transfer occurs, which enhances the chance for the ejection of the envelope during the spiral-in of the compact remnant of the primary. In short period binaries like 4U~1700-37, a merging of both components is rather more likely (cf. Podsiadlowski et al. 1995). The known black holes X-ray binaries which possibly all have high mass black hole companions (cf. Ergma \&\ van den Heuvel 1998, Baylin et al. 1998, Brown et al. 1999) have such low periods that none of them could have undergone a conservative evolution. The fact that non-conservative evolution occurs for all Case~C systems but not for all Case~A or~B systems is in agreement with the origin of high mass black holes in binaries as due to Case~C mass transfer (cf. Table~2), as proposed by Brown et al. (1999). The predominance of short periods in massive black hole binaries may be due to selection effects. On the other hand, neutron star binaries with much longer periods, consistent with conservative evolution, have been found (e.g. Wray~977/GX~301-2)). This is again consistent with the evolutionary scenario depicted above. It is important to note that Table~2 has been derived only for stars of about solar metallicity. The radius evolution of massive stars, and possibly also the Wolf-Rayet mass loss rates, depend strongly on metallicity. Also, we do not work out the relative frequencies with which the evolution through the various mass transfer channels occur. E.g., at solar metallicity, the Case~C evolution may be relatively rare (cf. Section~4), which could be the reason why only one Galactic high mass X-ray binary is know to contain a high mass black hole (Cyg~X-1), compared to two in the Large Magellanic Cloud. Furthermore, Wellstein et al. (1999) find that conservative evolution for Case~B systems is restricted to initial primary masses below $\sim 25\mso$, while Case~A systems are more likely to avoid contact for initial primary masses above $\sim 15\mso$. All this may change for different metallicities. To work this out has to be left to future investigations. \begin{acknowledgements} We are very grateful to Hans Bethe, Gerry Brown, Chris Fryer, Alexander Heger, Lex Kaper, Chang-Hwan Lee, and Stan Woosley for stimulating discussions and for the communication of results prior to publication. This work has been supported by the Deutsche Forschungsgemeinschaft through grants La~587/15-1 and 16-1. \end{acknowledgements}
\section{Introduction} \label{intro} From 1990 to 1995 the LEP $\ensuremath{\mathrm{e^+e^-}}$ storage ring was operated at centre of mass energies close to the Z mass, in the range $|\sqrt{s}-\mathrm{M_Z} | <3$~GeV. Most of the data have been recorded at the maximum of the resonance (120~pb$^{-1}$ per experiment) and about 2~GeV below and above (40~pb$^{-1}$ per experiment). The measurement of the hadronic and leptonic cross sections as well as the leptonic forward backward asymmetries performed with the \Aleph detector at these energies are presented here. The large statistic allows a precise measurement of these quantities which are then used to determine the Z lineshape parameters: the Z mass $\mathrm{M_Z}$, the Z width $\mathrm{\Gamma_Z}$, the total hadronic cross section at the pole $\mathrm{\sigma^0_{had}}$ and the ratio of hadronic to leptonic pole cross sections $\mathrm{R_l}=\mathrm{\sigma^0_{had}}/\mathrm{\sigma^0_{l}}$. Here we will give some details of the \Aleph experimental measurement of these quantities. A review of the whole LEP electroweak measurements and a discussion of the results as a test of the Standard Model can be found in another talk of this conference~\cite{Pinfold}. \section{Cross sections and leptonic Forward-Backward asymmetries measurement} \label{xs_afb} The cross section and asymmetries are determined for the s-channel process $\mathrm{e^+e^- \rightarrow Z,\gamma \rightarrow f\bar{f}}$. The cross section is derived from the number of selected events $\mathrm{N_{sel}}$ with \begin{equation} \mathrm{ \sigma_{f\bar{f}} = \frac{N_{sel}(1-f_{bkg})}{\epsilon} \frac{1}{\cal L}} \end{equation} where $\mathrm{f_{bkg}}$ is the fraction of background, $\mathrm{\epsilon}$ is the selection efficiency and ${\cal L}$ is the integrated luminosity. Note that in the case of $\ensuremath{\mathrm{e^+e^-}}$ final state the irreducible background originating from the exchange of $\gamma$ (Z) in the t-channel is subtracted from $\mathrm{N_{sel}}$ to obtain the s-channel cross section. The leptonic forward-backward asymmetry ($\mathrm{A_{FB}}$) is derived from a fit to the angular distribution \begin{equation} \mathrm{ \frac{d\sigma_{f\bar{f}}}{d\cos\theta^*}\propto 1+cos^2\theta^*+\frac{8}{3}A_{FB}cos\theta^*} \end{equation} where $\theta^*$ is the centre of mass scattering angle between the incoming $\mathrm{e^-}$ and the out-coming negative lepton. To achieve a good precision on the cross section a high efficiency and low background is necessary while the asymmetry is insensitive to the overall efficiency and to a background with the same asymmetry as the signal. This justifies the use of different leptonic selections for cross section and asymmetry measurement. \subsection{Hadronic cross sections} \label{xs_qq} Four million of $\mathrm{e^+e^-\rightarrow q\bar{q}}$ events have been recorded at the Z peak, leading to a statistical uncertainty of $0.05\%$. The systematic uncertainty has been reduced to the same level. The hadronic cross section measurement is based on two independent selections. The first selection is based on charged track properties while the second is based on calorimetric energy. Details of these selections can be found in previous publications~\cite{AlephLS}. These two measurements are in good agreement and are combined to obtain the final result. The systematic uncertainties of these selections are almost uncorrelated because they are mainly based on uncorrelated quantities therefore the combination of the 2 selections allows to reduce the systematic uncertainty. \begin{figure}[htpb] \vspace{-.6cm} \begin{center} \epsfxsize=15pc \vspace{-1cm} \epsfbox{etpc-ntpc.eps} \end{center} \caption{Distribution of charged track multiplicity ($\mathrm{N_{ch}}$) versus charged track energy ($\mathrm{E_{ch}}$). \label{TPChad}} \end{figure} Figure~\ref{TPChad} shows the distribution of charged multiplicity versus the charged track energy for signal and background. These variables are used to separate $\mathrm{e^+e^-\rightarrow q\bar{q}}$ events from background in the charged track selection. The most dangerous background in both selections is $\mathrm{\gamma\gamma}$ events because Monte Carlo prediction is not fully reliable. Therefore a method to determine this background from the data has been developed. This is achieved by exploiting the different $\sqrt{s}$ dependence of the resonant (signal) and non-resonant (background) contributions. The resultant systematic error ($0.04\%$) reflects the statistic of the data. \begin{table}[h] \vspace{-.4cm} \caption{Efficiency, background and systematic errors for the two hadronic selections at the peak point.} \begin{center} \begin{tabular}{|c|c|c|} \hline & { Charged tracks} & { Calorimeter}\\ \hline { Efficiency ($\%$) } & $97.48$ & $99.07$ \\ \hline { Background ($\%$)}& & \\ $\tau^+\tau^-$ & 0.32 & 0.44 \\ $\gamma\gamma$ & 0.26 & 0.16 \\ $e^+e^-$ & - & 0.08 \\ \hline { Systematic ($\%$) } & &\\ Detector simulation & 0.02 & 0.09\\ Hadronisation modelling & 0.06 & 0.03\\ $\tau^+\tau^-$ bkg & 0.03 & 0.05 \\ $\gamma\gamma$ bkg & 0.04 & 0.03 \\ $e^+e^-$ bkg & - & 0.03 \\ \hline Total syst. & 0.087 & 0.116\\ { Combined } & \multicolumn{2}{c|}{ 0.071}\\ \hline \multicolumn{3}{c}{ }\\ \multicolumn{3}{c}{ }\\ \end{tabular} \end{center} \label{qqsys} \end{table} The dominant systematic error in the calorimetric selection comes from the calibration of calorimeters ($0.09\%$) and, in the charged track selection from the determination of the acceptance ($0.06\%$). Table~\ref{qqsys} gives a breakdown of the efficiency, the background and the systematic uncertainties of both selections. \subsection{Leptonic cross sections} \label{xs_ll} The statistical uncertainty in the leptonic channel is of the order of $0.15\%$, The aim of the analysis is to reduce the systematic to less than $0.1\%$. Two analyses were developed for the measurement of the leptonic cross sections. The first one, referred to as {\it exclusive} is based on three independent selections each aimed at isolating one lepton flavour and still follows the general philosophy of the analysis procedures described earlier~\cite{AlephLS}. The second one is new and has been optimised for the measurement of $\mathrm{R_l}$, it is refered as {\it global} di-lepton selection. The results of these selections agree within the uncorrelated statistical error and have been combined for the final result. These selections are not independent since they make use of similar variables therefore their combination does not reduce the systematic uncertainty. We concentrate here on the {\it global} analysis. This selection takes advantage of the excellent particle identification capabilities (dE/dx, shower developement in the calorimeters and muon chamber information) and the high granularity of the \Aleph detector. First di-leptons are selected within the detector acceptance with an efficiency of $99.2\%$ and the background arising from $\mathrm{\gamma\gamma}$, $\mathrm{q\bar{q}}$ and cosmic events is reduced to the level of $0.2\%$. Then the lepton flavour separation is performed inside the di-lepton sample so that the systematic uncertainties are anti-correlated between 2 lepton species and that no additional uncertainty is introduced on $\mathrm{R_l}$. Table~\ref{globalsys} gives a breakdown of the systematic errors obtained with 1994 data. The background from $\mathrm{q\bar{q}}$ and $\mathrm{\gamma\gamma}$ events affects mainly the $\mathrm{\tau^+\tau^-}$ channel, therefore this channel is affected by bigger selection systematics than $\ensuremath{\mathrm{e^+e^-}}$ and $\mathrm{\mu^+\mu^-}$. As an example we consider the systematic errors related to the $\mathrm{\tau^+\tau^-}$ selection efficiency. This efficiency is measured on the data: $\mathrm{\tau^+\tau^-}$ events are selected using tight selection criteria to flag $\tau$-like hemispheres; with the sample of opposite hemispheres, {\it artificial} $\mathrm{\tau^+\tau^-}$ events are constructed by associating two back-to-back such hemispheres and the selection cuts are applied. In order to assess the validity of the method and to correct for possible bias of this method, two different Monte Carlo reference samples are used. On the first sample the same procedure of artificial $\mathrm{\tau^+\tau^-}$ events is applied, on the second one the selection cuts are applied directly. The uncertainty on the efficiency measured with this method is dominated by the statistic of the artificial events used in the data. This method is applied in order to measure the inefficiency arising from $\mathrm{q\bar{q}}$ cuts and in the flavour separation. The dominant systematic in $\ensuremath{\mathrm{e^+e^-}}$ channel arises from t-channel subtraction and is given by the theoretical uncertainty~\cite{alistar} on the t-channel contribution to the cross section. This leptonic cross section measurement contributes to a systematic uncertainty on $\mathrm{R_l}$ of $0.08\%$. \begin{table}[t] \caption{ Systematic uncertainties in percent of dilepton cross sections for peak 1994 data. Correlations between lepton flavours are taken into account in the $\mathrm{l^+l^-}$ column. \label{globalsys}} \begin{center} \begin{tabular}{|l|c|c|c||c|} \hline & $\ensuremath{\mathrm{e^+e^-}}$ & $\mathrm{\mu^+\mu^-}$ & $\mathrm{\tau^+\tau^-}$ & $\mathrm{l^+l^-}$ \\ \hline \multicolumn{5}{|c|}{ Global selection}\\ \hline Tracking efficiency& 0.05 & 0.03 & 0.03&0.04 \\ \hline Angles measurement (*)&0.02 & 0.01 & 0.01 & 0.02 \\ \hline ISR and FSR simulation (*)& 0.03 & 0.03 &0.03 & 0.03 \\ \hline $\mathrm{\gamma\gamma}$ cuts (*)& 0.02 & - & 0.05 & 0.02 \\ \hline $\mathrm{q\bar{q}}$ cuts& - & - & 0.11 & 0.04 \\ \hline $\mathrm{\gamma\gamma}$ background (*) & - & - & 0.02 & -\\ \hline $\mathrm{q\bar{q}}$ background(*) & - & - & 0.04 & 0.01 \\ \hline \multicolumn{5}{|c|}{Flavour separation}\\ \hline $\mathrm{\mu^+\mu^-}$/$\mathrm{\tau^+\tau^-}$ & - & 0.03 & 0.03 & - \\ \hline $\ensuremath{\mathrm{e^+e^-}}$/$\mathrm{\tau^+\tau^-}$& \multicolumn{4}{|c|}{ }\\ $cos\theta^*<0.7$& 0.08 & - & 0.07 & 0.01\\ $cos\theta^*\geq0.7$& - & - & 0.06 & 0.02\\ \hline \multicolumn{5}{|c|}{t-channel subtraction}\\ \hline (*) & 0.11&-&-& 0.04 \\ \hline \multicolumn{5}{|c|}{Monte Carlo statistic}\\ \hline & 0.05 & 0.06 & 0.07 & 0.04 \\ \hline \hline Total & 0.16 & 0.08 &0.19 & 0.09 \\ \hline \multicolumn{5}{c}{(*)\footnotesize uncertainties completely correlated among all energy points.} \end{tabular} \end{center} \end{table} \subsection{Leptonic Forward-Backward asymmetries} \label{Afb} The measurement of the asymmetries is dominated by the statistical uncertainty equal to 0.0015. Special muon and tau selections have been designed for the asymmetry measurement while the $\ensuremath{\mathrm{e^+e^-}}$ {\it exclusive} selection is used for the $\ensuremath{\mathrm{e^+e^-}}$ asymmetry measurement. The $\ensuremath{\mathrm{e^+e^-}}$ angular distribution needs to be corrected for efficiency before subtracting the t-channel and therefore relies on Monte Carlo while $\mathrm{\mu^+\mu^-}$ and $\mathrm{\tau^+\tau^-}$ angular distributions do not need to be corrected with Monte Carlo since the selections are designed so that the efficiency is symmetric. Because of the $\gamma-Z$ interference, the asymmetry varies rapidly with the centre of mass energy $\sqrt{s'}$ around the Z mass. Cuts on energy induce a dependence of the efficiency with $\sqrt{s'}$ and therefore could introduce a bias in the measurement since the efficiency would no more be symmetric. To minimise these effects the selections are mainly based on particle identification instead of kinematic variables. \begin{figure}[htpb] \begin{center} \epsfxsize=17pc \epsfbox{leptpid_ewpap.eps} \end{center} \caption{Identification efficiencies for a) electrons, b) muons and c) pions as a function of the momentum to the beam energy ratio. In d) the probability for a pion to be misidentified as an electron (squares) or a muon (circles) is shown. \label{leptid}} \end{figure} Figure \ref{leptid} shows the particle identification efficiency as a function of energy. The asymmetry is extracted by performing a maximum likelihood fit to the differential cross section. The dominant systematic uncertainty arises from t-channel subtraction in the Bhabha channel (0.0013 on $\mathrm{A_{FB}^{0,e}}$) and other systematic errors are smaller than 0.0005. \begin{figure}[htpb] \vspace{-.8cm} \hspace{-2cm} \begin{minipage}[b]{\textheight} \begin{center} \epsfxsize=19pc \epsfbox{lineshape.eps} \end{center} \caption{Measurement of cross sections. The inserts show enlarged views of the peak region. \label{lineshape}} \end{minipage} \vspace{-1cm} \hspace{-2cm} \begin{minipage}[b]{\textheight} \begin{center} \epsfxsize=19pc \epsfbox{asymmetries.eps} \end{center} \caption{Measurement of the asymmetries. The inserts show enlarged views of the peak region. \label{asymmetries}} \end{minipage} \end{figure} \section{Results} \label{results} Figures~\ref{lineshape} and \ref{asymmetries} show the measured cross sections and asymmetries. The Z lineshape parameters are fitted to these measurements with the latest version of ZFITTER~\cite{ZFITTER}. The error matrix used in the $\chi^2$ fit includes the experimental statistic and systematic uncertainties, the LEP beam energy measurement uncertainty and the theoretical uncertainties arising from the small angle Bhabha cross section in the luminosity determination and the t-channel contribution to wide angle Bhabha events. The results are shown in Table~\ref{res_tab}. \begin{table}[t] \caption{Result of the fit to experimental measurement of cross sections and leptonic forward-backward asymmetries. The total error is splitted into statistical and systematic errors, theoretical error on luminosity and on t-channel and LEP beam energy measurement error.\label{res_tab}} \begin{center} \begin{tabular}{|c|c|r|r|r|r|r|} \hline & value & stat & exp & { lumi} & { t-ch} & { $E_{beam}$} \\ \hline $\mathrm{M_Z}$ & 91.1883$\pm$0.0031 & 0.0024 & 0.0002 & & & { 0.0017} \\ \hline $\mathrm{\Gamma_Z}$ & 2.4953$\pm$0.0043 & 0.0038 & 0.0009 & & &{ 0.0013}\\ \hline $\mathrm{\sigma^0_{had}}$ & 41.557$\pm$0.058 & 0.030 & 0.026 & {0.025} & &{ 0.011}\\ \hline $\mathrm{R_el}$ & 20.677$\pm$0.075 & 0.062 & 0.033 & & { 0.025} & { 0.013}\\ \hline $\mathrm{R_\mu}$ & 20.802$\pm$0.056 & 0.053 & 0.021 & & & {0.006}\\ \hline $\mathrm{R_\tau}$ & 20.710$\pm$0.062 & 0.054 & 0.033 & & & {0.006}\\ \hline $\mathrm{A_{FB}^{0,e}}$ & 0.0189$\pm$0.0034 & 0.0031 & 0.0006 & & { 0.0013}& { 0.0002}\\ \hline $\mathrm{A_{FB}^{0,\mu}}$ & 0.0171$\pm$0.0024 & 0.0024 & 0.0005 & & & { 0.0002}\\ \hline $\mathrm{A_{FB}^{0,\tau}}$ & 0.0169$\pm$0.0028 & 0.0026 & 0.0011 & & & { 0.0002}\\ \hline \hline $\mathrm{R_l}$ & 20.728$\pm$0.039 & 0.033 & 0.020 & & { 0.005} & { 0.002}\\ \hline $\mathrm{A_{FB}^{0,l}}$ & 0.0173$\pm$0.0016 & 0.0015 & 0.0004 & & { 0.0002} & { 0.0001}\\ \hline \end{tabular} \end{center} \end{table} The value of the Z couplings to charged leptons $|g_V|$ and $|g_A|$ can be derived from these parameters. The experimental measurement is shown in Figure \ref{gvga} with the Standard Model prediction. The data favor a light Higgs. The value of $\alpha_s$ can also be extracted from $\mathrm{R_l}$, $\mathrm{\Gamma_Z}$ and $\mathrm{\sigma^0_{had}}$: \begin{equation} \mathrm{\alpha_s = 0.115 \pm 0.004_{exp} \pm 0.002_{QCD} } \end{equation} where the first error is experimental and the second reflects uncertainties on the QCD part of the theoretical prediction~\cite{ALPHAS}. Here the Higgs mass has been fixed to 150 GeV and the dependence of $\alpha_s$ with $\mathrm{M_H}$ can be approximately parametrised by $\mathrm{\alpha_s(M_H)=\alpha_s(M_H=150 GeV)\times(1+0.02\times ln(M_H/150))}$ where $\mathrm{M_H}$ is expressed in GeV. \section{Conclusion} \label{conclusion} The high statistic accumulated by \Aleph during LEP 1 running allows to measure the hadronic and leptonic cross sections and the leptonic forward-backward asymmetries with statistical and systematic precision of the order of 1 permil. These measurements are turned into precise determination of the Z boson properties and constraints on the Standard Model parameters. \begin{figure}[t] \begin{center} \epsfxsize=15pc \epsfbox{va.eps} \end{center} \caption{Effective lepton couplings. Shown are the one-$\sigma$ contours. The shaded area indicates the Standard Model expectation for $\mathrm{M_t}=174\pm5$~GeV/$c^2$ and $90< \mathrm{M_H} GeV/c^2<1000$; the vertical fat arrow shows the change if the electromagnetic coupling constant is varied within its error. The sign of the couplings were assigned in agreement with $\tau$ polarisation measurements and neutrino data scattering. \label{gvga}} \end{figure} \section*{Acknowledgments} I would like to thank D. Schlatter for his help in preparing this talk. I also would like to thank the organisers of the Lake Louise Winter Institute conference for the interesting program and the nice atmosphere at the conference.
\section{Introduction} \subsection{Structure of the work} After background and setting of a problem I present the short introduction to abstract quantum computations in the section 3. In the section 4 linear differential equations are applied to a tight analysis of the famous Grover algorithm of the fast quantum search. The section 5 is the key. Here a parallel quantum algorithm for repeated search is defined and studied by means of differential equations. In the section 6 we briefly run through a parallel algorithm for iterated search with simultaneous queries of all oracles. The abstract includes the other substantiation of the algorithm. \subsection{Background} The most promising quantum mechanical application to the algorithm theory is associated with the fundamental algorithm of exhaustive search or finding a solution of equation $f(x)=1$ for a Boolean function $f$. In 1996 Lov Grover in the work \cite{Gr1} showed how quantum computer can solve this equation for the case of unique solution in a time $O(\sqrt{N})$ where $N$ is the number of all possible values for $x$, whereas every probabilistic classical algorithm requires a time $O(N)$. At about the same time in the work \cite{BBBV} it was shown that there are not substantially faster algorithm for this problem. Later a tight estimation for the time of Grover's algorithm as $\frac{\pi\sqrt{N}}{4}$ with the probability of error about $1/N$ was established in the work \cite{BBHT}. A further development of the fast quantum search can be found in the works \cite{DH}, \cite{FG}, \cite{Jo}, \cite{CGW}, \cite{Ro}, \cite{BBBGDL}, \cite{JMH}, \cite{H}. The earlier patterns of quantum speedup was constructed by P. Shor (look at the work \cite{Sh}). There are the algorithms finding a factorization of an integer and a discrete logarithm. Algorithms of such a sort are presented in a lot of works (look for example at \cite{DJ}, \cite{Gr2}, \cite{Ki1}, \cite{Si}, \cite{St}, \cite{TS}, etc.). The classical computations admitting quantum speedup are rare exclusions from all classical computations in the following sense. Denote all words of a length $n$ in the alphabet $\{ 0,1\}$ by $\{ 0,1\}^n$. We can represent a general form of classical computation as $T$ iterated applications of some oracle $g:\ \{ 0,1\}^n \longrightarrow \{ 0,1\}^n$: \begin{equation} x\longrightarrow g(x)\longrightarrow g(g(x))\longrightarrow\ldots\longrightarrow g(g(\ldots (g(x)))). \label{1} \end{equation} In the work \cite{Oz1} it is shown that if $T=O(N^{\frac{1}{7+\epsilon}} ),\ \epsilon>0$, then for the bulk of all $g$ such computation has not any quantum speedup. Similar results for search problems were obtained in the works \cite{BBBV}, \cite{BBHT}, \cite{Oz2}, \cite{Za}. A lower bound as $O(N)$ for a time of quantum computations of functions with functional argument $F:\ \{ f\}\longrightarrow\{ 0,1\}$ was found in the work \cite{BBCMW}. Here $f$ are Boolean functions on domain of cardinality $N$. At the same time using a memory of $O(N)$ qubits it is possible to compute such functions in a time $N/2$ (look at \cite{vD}). This brings up the question: what a general type of classical computations of the form (1) admits a quantum speedup beyond any possible speedup of $g$? As follows from the work \cite{Oz1} for the bulk of functions $g$ this speedup can result only from a parallel application of $g$. About other aspects of quantum evolutions look also at \cite{ML}, \cite{Ho}, \cite{Pe}. \section{Setting of the problem} Consider the following situation. We want to gather a mosaic from scattered stones in a rectangular list with the corresponding picture. Each stone is of a unique form. We can gather this mosaic layer by layer and use a simple search among stones still scattered to fill any layer basing on the previous one. Then we in fact fulfill an iterated search classically, because to find the stones for the following layer we must already have the previous one filled. We formalize this as the special type of iterated algorithms: iterated search (IS). Suppose we have a sequence $S_1 , S_2 ,\ldots , S_k$ of similar search problems where $S_i$ is to find a unique solution $x_i^0$ of equation $f_i (x_i )=1$ where a Boolean function $f_i$ is accessible iff we know all $x_j^0 ,\ \ j<i,$ $|x_i |=n,\ \ N=2^n ,\ \ k\ll N$, $|x|$ denotes a length of word $x$. The aim is to discover $x_k^0$, $k\geq 2$. In view of the result of the work \cite{BBHT} cited above sequential applications of Grover's search for $x_1^0 ,x_2^0 ,\ldots , x_k^0$ give an answer in the time $\frac{k\pi\sqrt{N}}{4}$ with error probability about $k/N$. To do this we must have all oracles $f_i,\ \ i=1,2, \ldots , k$, where the dependence $f_i$ of all $x_j ,\ \ j<i$ can be included to $f_i$. So we can assume that $f_i$ has the form $f_i (x_1 , x_2 ,\ldots ,x_i )$ and each equality $f_i (x_1 , \ldots , x_i )=1$ has the unique solution $x_1^0 , x_2^0 , \ldots , x_i^0 ,\ \ i=1,2,\ldots , k.$ Considering all oracles $f_i$ as physical devices which can not be cloned we assume that they are in our disposal at the same time, so we can apply them simultaneously. In such application advantage is taken of interference between the results of their actions. This results in a speedup of computation comparatively with the sequential mode. Why this speedup can arise? It arises because of a leak of amplitude in each step of sequential search. A leak of amplitude issues from that an amplitude of $x_i^0$ in search number $i$ increases step by step in course of Grover's search, after few first $l$ steps it becomes approximately $\frac{2l+1}{\sqrt{N}}$, when amplitudes of others $x_i \neq x_i^0$ decrease. This prevailing of the amplitude corresponding to $x_i^0$ (a leak of amplitude) can be immediately used for the next $i+1$-th search. We shall show how this effect can be used to solve the problem of iterated search in a time $O(\frac{k\pi\sqrt{N}}{4\sqrt{2}})$ which is $\sqrt{2}$ times as fast as by $k$ sequential applications of Grover's algorithm. Thus we shall have a modification of the fast quantum search - the parallel quantum algorithm for iterated search which will be described later. In this article we take up mostly the particular case $k=2$ of IS, we name this problem repeated search (RS). RS-problem is connected with the known problem of structured search (SS). The problem of structured search is to find a unique solution $x_0 ,y_0$ of $f(x,y)=1$ provided we have a function $g$ whose support $\{ x\ |\ g(x)=1\}$ of cardinality $M$ contains $x_0$. RS-problem is a particular case of SS when $M=1$. The case $1\ll M\ll N$ was investigated by Farhi and Gutmann in the work \cite{FG}. They have found quantum algorithm for this case with a time complexity $O(\sqrt{MN})$, and also they wrote that the best known strategy for the case $M=1$ is the sequential application of Grover algorithm. In the present paper it is shown how this evident strategy can be improved in constant factor $\sqrt{2}$. At last note that our approach differs from the work \cite{FG}. Farhi and Gutmann used only algebraic properties of Grover's algorithm whereas for RS-problem we need to work with an evolution of amplitudes in computation. \section{ Quantum computations and differential equations} After early studies of R. Feynman (\cite{Fe}), P. Benioff (\cite{Be}) and D. Deutsch (\cite{De}) numerous approaches to quantum computations have appeared (look at \cite{BV}, \cite{Wa}, \cite{Ya}, \cite{Ll1}). Leaving aside the problem of decoherence and quantum codes (look at the articles \cite{AB} , \cite{CLSZ}, \cite{Ki2}, \cite{Pr} ) we shall regard ideal computations in closed systems. We use the abstract model of quantum computer independent of the formalism of classical algorithm theory. This model consists of two parts: classical and quantum (look at the picture 1). A state $C$ of {\it Classical part} consists of the following objects. 1) Registers with labels corresponding to transformations $U_{i_j}$ of the finite set $\{ U_i \}$ of elementary unitary transformations with no more than 3 qubits each. (Strictly speaking transformations on two qubits would suffice: look for example at the work \cite{DiV}). Moreover, as follows from the works \cite{Ll2}, \cite{BBCDMSSSW} there is a variety of possible choices of the set $\{ U_{i_j} \}$. 2) Pointers aimed from these registers to as many qubits from the quantum part as there are arguments of the corresponding unitary transformation. Here each qubit is involved in exactly one transformation. 3) Registers of an end of computation and of a query: $e(C)$ and ${\rm qu} (C)$ respectively. A {\it Quantum part} is a tape partitioned into cells with one qubit each. Every qubit has two basic states $|0\rangle ,\ |1\rangle$, so its quantum state $\lambda |0\rangle +\mu |1\rangle$, $|\lambda |^2 +|\mu |^2 =1$, belongs to the curcle of radius 1 in two dimensional Hilbert space ${\rm C}^2$. If $n$ is a length of tape, all states of the tape belong to the tensor product ${\cal H} =\underbrace{{\rm C}^2 \bigotimes {\rm C}^2 \bigotimes\ldots \bigotimes {\rm C}^2} _{n}$ of spaces, corresponding to all qubits that is ${\cal H}={\rm C}^{2^n}$. Each state of quantum part is a superposition $\chi =\sum\limits_{i=0}^{N-1} \lambda_i e_i$ of basic states $e_0 ,\ldots , e_{N-1}$ with complex amplitudes $\lambda_i$ where $\sum\limits_{i=0}^{N-1} |\lambda_i |^2 =1$, $N=2^n$. We can assume that all $e_i \in\{ 0,1\}^n$. An {\it Observation} of this state $\chi$ is a random variable which takes each value $e_i$ with the probability $|\lambda_i |^2$. A {\it Working transformation} of quantum part corresponding to a fixed state of classical part has the form $U_{i_1} \bigotimes U_{i_2} \bigotimes\ldots \bigotimes U_{i_k}$ where each $U_{i_j}$ acts on qubits which the corresponding pointer aims to. Let $f_1 , \ldots , f_l$ be functions of the form $\{ 0,1\}^n \longrightarrow \{ 0,1\}^m$ and for each $i=1,2,\ldots , l$ there are special places in the quantum tape reserved for an argument $a_i$ of $f_i$ (query) and for a value of $f_i$ (answer). Denote by $b_i$ an initial contents of the place for answer. A {\it Query transformation} ${\rm Qu} _{\bar f}$ is ${\rm Qu} _{f_1} \bigotimes {\rm Qu} _{f_2} \bigotimes\ldots \bigotimes {\rm Qu} _{f_l}$ where for each $i=1,\ldots , l$ \newline ${\rm Qu} _{f_i} \ |a_i ,b_i \rangle \longrightarrow |a_i , b_i \bigoplus f_i (a_i ) \rangle$, $ \bigoplus$ is bitwise addition modulo 2. We name these functions ${\rm Qu}_{f_i}$ oracles. A {\it Quantum algorithm} is an algorithm determining evolution of the classical part: \begin{equation} C_0 \longrightarrow C_1 \longrightarrow \ldots \longrightarrow C_T \label{2} \end{equation} (in particular it determines a number $T$). A classical part plays the role of controller for quantum part and determines its evolution (look at the picture 2). A {\it Quantum computation} consists of two sequences: (2) and \begin{equation} Q_0 \longrightarrow Q_1 \longrightarrow \ldots \longrightarrow Q_T \label{3} \end{equation} where for each $i=0,1,\ldots ,T-1$ $e(C_i )=0;$ $e(C_T )=1$ and every passage $Q_i \longrightarrow Q_{i+1}$ is: - a working transformation, corresponding to $C_i$, if ${\rm qu} (C_i )=0,\ e(C_i )=0$, - a query transformation ${\rm Qu}_{\bar f}$, if ${\rm qu} (C_i )=1,\ e(C_i )=0$. A {\it result} of this quantum computation is a contents of first $n_0$ qubits of quantum tape after the observation of final state $Q_T$. An initial state $(C_0 , Q_0 )$ of the computer is obtained from an input data $\bar x$ by some routine procedure. A {\it time} of computation (2),(3) is a number of query transformations (queries) in it. We see that in this model some oracles may be called simultaneously which causes interference between results of their actions. We shall prove that such interference can speed up computations in case of repeated search. From a physical standpoint systems of linear differential equations is a natural tool for the study of quantum computation. Wave function $\psi$ which determins an evolution of quantum computer satisfies Shr\"odinger equation $\frac{\partial \psi}{\partial t} =iH\psi$, where $H$ is Hamiltonian within a real factor. An evolution of $\psi$ is determined by unitary operator $U(t):$ $\psi (t)=U\psi (0)$ which satisfies the equation $\dot U=iHU$. This equation is a prototype of all systems of differential equations studied below. We need only to choose Hamiltonian so that the amplitude of target state peaks in a point $t_{quant}$ which is less than the time of the best classical computation. In rare cases quantum parallelism makes it possible. Some other aspects of parallelism in computing may be found in \cite{BM}, \cite{BO}, \cite{LMS}, \cite{Wo}. \section{ An exact description of simple quantum search by differential equations} \subsection{Notations} Assume the notations of Dirac where a vector $\bar a\in {\rm C}^m$ as a column of coordinates is denoted by $|\bar a \rangle$. A row obtained from $|\bar a\rangle$ by the transposition and complex conjugation is denoted by $\langle \bar a |$. A dot product of $\bar a,\bar b\in {\rm C}$ will be $\langle \bar a|\bar b\rangle$. A result of application of operator $A$ to a vector $|\bar a\rangle$ is denoted by $A|\bar a \rangle$. For every transformations $A,\ B$ of the forms ${\cal L}_1 \longrightarrow {\cal L}_2 ,\ {\cal L}_2 \longrightarrow {\cal L}_3$ we denote by $AB$ their composition which acts from right to left such that $AB(x) =A(B(x))$. Given vectors $a\in{\cal L} ,\ b\in {\cal M}$ from linear spaces $\cal L$, $\cal M$ the state $|a\rangle\bigotimes |b\rangle \in {\cal L}\bigotimes {\cal M}$ is denoted by $|a,b\rangle$. For a function $F:\ \ F|X,Y\rangle =|X,\phi (Y) \rangle$ we denote by $F|_Y$ its restriction on $Y:\ \ F|_Y |Y\rangle =|\phi(Y)\rangle$. Let $f$ be a function of the form $A\longrightarrow A$. We define an $i$-th iteration of $f$: $f^{\{i\}}$ by the following induction on $i$. Basis: $f^{\{1\}} =f$. Step: $f^{\{i+1\}} =ff^{\{i\}}$. \subsection{ Grover's quantum algorithm for simple search and its implementation in our model} Grover's algorithm for finding a unique solution $x_0$ of equation $f(x)=1$ for a Boolean $f$ is sequential applications of the following unitary transformation: $G= -WR_0 WR_t$ to the initial state $\chi_0 = \frac{1}{\sqrt{N}}\sum\limits_{i=0}^{N-1} |e_i \rangle$ where Walsh-Hadamard transformation is $W=\underbrace{J\bigotimes\ldots\bigotimes J}_{n}$, $$ J=\left(\begin{array}{cc} 1/\sqrt 2\ &1/\sqrt 2 \\ 1/\sqrt 2 \ &-1/\sqrt 2 \end{array}\right), \ \ R_0 |e\rangle = \left\{ \begin{array}{cc} |e\rangle ,\ &\mbox{if } e\neq \bar 0,\\ -|0\rangle ,\ &\mbox{if } e=\bar 0, \end{array} \right.\ \ R_t |e\rangle = \left\{ \begin{array}{cc} |e\rangle ,\ &\mbox{if } e\neq x_0,\\ -|x_0 \rangle ,\ &\mbox{if } e=x_0 .\end{array} \right. $$ It is easily seen that $W$ can be implemented on our model of quantum computer. To implement $R_t$ it is sufficient to apply ${\rm Qu}_f$ to the state $\frac{1}{\sqrt{N}} \sum\limits_{i=0}^{N-1} |e_i \rangle \bigotimes \frac{|0\rangle -|1\rangle}{\sqrt{2}}$, where the last qubit is the place for oracle's answer. To implement $R_0$ we need $n+1$ ancillary qubits initialized by $0$. Consider some function $\phi$ acting on three qubits: $|\mbox{main, ancilla, }res\rangle$ as follows. $$ \begin{array}{cc} |000\rangle\ &\longrightarrow\ |000\rangle\\ |100\rangle\ &\longrightarrow\ |101\rangle\\ |001\rangle\ &\longrightarrow\ |001\rangle\\ |101\rangle\ & \longrightarrow\ |111\rangle . \end{array} $$ Apply $\phi$ sequentially after each step moving pointers to right on one qubit in the main and ancillary areas (look at the picture 3). This makes $res =1$ iff at least one of the main qubits is 1. Then inverse the phase of $0$ in the qubit $res$ and fulfill all reverse transformations with $\phi$ in the reverse order restoring the initial states of ancillary qubits. Let $\chi_i =a_i \sum\limits_{e' \neq e} |e'\rangle +b_i |e\rangle ,$ $\chi_{i+1} =G \chi_i$, $e$ is a state of quantum part, corresponding to a target word $x_0$. The difference between $x_0$ and $e$ is that $e$ contains also ancillary qubits whose values will be restored after each step of computation (it can be simply traced in what follows). The main property of Grover's transformation may be represented by the following equations (look at \cite{Gr1}, \cite{BBHT} ). \begin{equation} \left\{\begin{array}{cl} b_{i+1} &= (1-\frac{2}{N} )b_i +2(1-\frac{1}{N} )a_i ,\\ a_{i+1} &=-\frac{2}{N} b_i +(1-\frac{2}{N} )a_i .\\ \end{array}\right. \label{4} \end{equation} \subsection{ The passage to the system of differential equations } The system (4) can be rewritten in the following form \begin{equation} \left\{\begin{array}{cl} b_{i+1} -b_i &= -\frac{2}{N} b_i +2(1-\frac{1}{N} )a_i ,\\ a_{i+1} -a_i &=-\frac{2}{N} b_i -\frac{2}{N} a_i ,\\ \end{array}\right. \label{5} \end{equation} where the initial condition of quantum part gives $a_0 =b_0 =1/\sqrt{N}$. This system (5) with the initial condition is the system of difference equations approximating the following system of linear differential equations: \begin{equation} \left\{\begin{array}{cl} \dot b\delta &= -\frac{2}{N} b +2(1-\frac{1}{N} )a,\\ \dot a\delta &=-\frac{2}{N} b -\frac{2}{N} a,\\ \end{array}\right. \label{6} \end{equation} with two unknown functions $a(t), b(t)$, constant $\delta >0$ and the initial condition $a(0)=b(0)=\/\sqrt{N}$, where $a_i ,b_i$ approximate $a(i\delta ), b(i\delta )$; $\delta$ is a step. A difference between solutions of (5) and (6) on a segment of the form $t\in [0,O(\delta\sqrt{N} )]$ is $O(\sqrt{N} \delta ^2 )$, hence the error of this approximation may be done as small as required by varying $\delta$ (an integral part of number $x$ is denoted by $[x]$). Solving (6) we obtain $$ \ddot b +\frac{4}{\delta^2 N} b+O(\frac{b}{\delta^2 N^2} )+\ddot b O( \frac{1}{N} )+\dot b O(\frac{1}{N\delta })=0. $$ Hence in within $O(\frac{1}{\sqrt{N}})$ a solution $b$ of (6) can be approximated by a solution of equation \begin{equation} \ddot b +\frac{4}{\delta^2 N} b=0 \label{7} \end{equation} with the initial conditions $b(0)=\frac{1}{\sqrt{N}},$ $\frac{1}{2} (\dot b (0)\delta +\frac{2}{N} b(0))=\frac{1}{\sqrt{N}}$ on the segment $[0,2/\omega ]$, where $\omega = 2/\delta \sqrt{N}$. The required solution of (7) with this accuracy is $b=\sin (\omega t)+ \frac{1}{\sqrt{N}}\cos (\omega t)$, and the maximum of amplitude 1 is in the point $t_0 = \frac{\pi\sqrt{N}}{4}\delta -\frac{\delta}{2}$. Then $\left[\frac{t_0}{\delta}\right] =\left[\frac{\pi\sqrt{N}}{4}\right] \stackrel{+}{-} 1$ recurrent steps (4) are necessary and sufficient to achieve this value of $b$ with this accuracy Thus we obtain that the accuracy $O(\frac{1}{\sqrt{N}})$ is reached in $\left[\frac{\pi\sqrt{N}}{4}\right]$ Grover's iterations. In the work \cite{BBHT} its authors obtained an exact solution of (4) and thus proved that in fact a probability of error is even about $1/N$. The approximation of amplitude's evolution by systems of differential equations is more universal method. For example it makes possible to handle with the more involved case of parallel algorithm for RS which is the subject of next section. \section{Parallel algorithm for the repeated quantum search} \subsection{ Definitions and result} Let $u,\ x,\ y$ be variables with values from three different copies of ${\cal H}_0 ={\rm C}^N$, $a=a_1 \bigotimes a_2 \in {\rm C}^4$, where $a_1 =a_2 =\frac{1}{\sqrt{2}} (|0\rangle -|1\rangle )$. We assume the notations $f_1 (x),\ f_2 (x,y)$ for two oracles in the repeated quantum search and let $e_1, e_2$ be such values for $x,y$ which represents unique solutions of equations $f_1 =1,\ f_2 =1$. We denote the corresponding states of quantum tape by the same letters. Put ${\cal H}={\cal H}_0 \bigotimes {\cal H}_0 \bigotimes {\cal H}_0 \bigotimes {\rm C}^4$. Let $$ \begin{array}{rl} F_1 |u,x,y,a\rangle &=|u,x,y,a_1 \bigoplus f_1 (u),a_2 \rangle ,\\ F_2 |u,x,y,a\rangle &=|u,x,y,a_1 ,a_2 \bigoplus f_2 (x,y)\rangle ,\\ P|u,x,y,a\rangle &=|u\bigoplus x,x,y,a\rangle. \end{array} $$ Then $$ F_1 |u,x,y,a\rangle =\left\{ \begin{array}{cc} |u,x,y,a\rangle ,\ &\mbox{if } u\neq e_1 ,\\ -|u,x,y,a\rangle ,\ &\mbox{if } u=e_1 ; \end{array} \right. $$ $$ F_2 |u,x,y,a\rangle =\left\{ \begin{array}{cc} |u,x,y,a\rangle ,\ &\mbox{if } |x,y\rangle\neq |e_1 ,e_2 \rangle ,\\ -|u,x,y,a\rangle ,\ &\mbox{if } |x,y\rangle =|e_1 ,e_2 \rangle ; \end{array} \right. $$ Define the following auxiliary unitary transformations on $\cal H$: ${\cal R}_0 =I\bigotimes R_{0x} \bigotimes R_{0y} \bigotimes I;\ \ {\cal W} =I\bigotimes W_x \bigotimes W_y \bigotimes I;\ \ {\cal F} =P(F_1 \mid_{u,a_1} \bigotimes F_2 \mid_{x,y,a_2} )P, $ where the lower indices $x,y$ point the corresponding area of application for Walsh-Hadamard transformations and rotations of the phase of $0$, $I$ denotes the identity. The key unitary transformation of parallel algorithm for RS is \begin{equation} Z={\cal WR}_0 {\cal WF} . \label{8} \end{equation} {\it The parallel algorithm for RS} is the sequential applications of $Z$ beginning with the initial state $$ \chi_0 =| \bar 0\rangle \bigotimes\frac{1}{\sqrt{N}}\sum\limits_{i=0}^{N-1} |e_i \rangle\bigotimes\frac{1}{\sqrt{N}}\sum\limits_{i=0}^{N-1} |e_i \rangle \bigotimes a $$ $\left[\frac{\pi\sqrt{N}}{2\sqrt{2}}\right]$ times. \newtheorem{Theorem}{Theorem} \begin{Theorem} Let $t=\left[\frac{\pi\sqrt{N}}{2\sqrt{2}}\right]$. Then the observation of $Z^{\{t\}} (\chi_0 )$ gives $x=e_1,\ y=e_2$ with probability of error $O(\frac{1}{\sqrt{N}} )$. \end{Theorem} \subsection{ An advantage of parallel quantum algorithm} It follows from the definition of $Z$ that oracles $F_1$ and $F_2$ for functions $f_1,\ f_2$ work simultaneously in parallel hence the algorithm requires approximately $\left[ \frac{\pi\sqrt{N}}{2\sqrt{2}} \right]$ simultaneous queries to obtain a result, when the sequential application of simple quantum searches with $f_1$ and then with $f_2$ requires $\left[ \frac{\pi\sqrt{N}}{2} \right]$ time steps to obtain a result with the same probability. Note that for a simple search constant factor $\frac{\pi\sqrt{N}}{4}$ can not be essentially improved (look at \cite{BBHT}). Suppose that every query is fulfilled by a physical device (oracle) of the peculiar type corresponding to a form of query. A set with a minimum of oracles which is necessary for the solution of RS consists of one oracle for $f_1$ and one for $f_2$. With these oracles we can run them simultaneously in the parallel algorithm and obtain a result $\sqrt{2}$ times faster than by the sequential search. But if we have {\it two copies} of each oracle it is possible to achieve the same performance by sequential search if we divide the whole domain $\{ 0,1\}^n$ into two equal parts of $N/2$ elements each and apply a simple quantum search at first with two copies of $f_1$-oracle - one for each part, then, having $e_1$, with two copies of $f_2$-oracle. But this last way is expensive if every copy of oracle has a large cost, or impossible at all if every oracle is unique, say issues from a natural phenomenon. Just in this case of minimal possible set of oracles $f_1$, $f_2$ the application of parallel quantum algorithm for RS gives the increasing of performance in $\sqrt{2}$ times. This speedup can be also obtained for IS if we apply this algorithm sequentially for the pairs $f_i , f_{i+1} ,\ i=1,2,\ldots , k-1.$ The resulting error probability will be $O(k/\sqrt{N})$. The remaining part of this work is devoted to the proof of Theorem and perspectives of this approach. \subsection{ A primary analysis of parallel algorithm for RS} Note that each of $W_y ,\ R_{0y}$ commutes with $W_x ,\ R_{0x} ,\ P, \ F_1$; $P$ commutes with $F_2$, hence $Z$ can be represented in the form $$ Z=-[(I\bigotimes W_x R_{0x} W_x \bigotimes I)PF_1 P][-(I\bigotimes ( W_y R_{0y} W_y )\bigotimes I) F_2 ], $$ or in the form \begin{equation} Z=\{ -W_x R_{0x} W_x {\cal F}_1 \}\{ -W_y R_{0y} W_y F_2 \} , \label{9} \end{equation} where $$ {\cal F}_1 |u,x,y,a\rangle =\left\{ \begin{array}{cl} |u,x,y,a\rangle \ \ &\mbox{if } x\neq e_1 ,\\ -|u,x,y,a\rangle \ \ &\mbox{if } x=e_1 . \end{array} \right. $$ The form (9) is exactly the repetition of Grover's transformations with oracles $F_2 ,\ {\cal F}_1$ in this order, hence we can apply the formulas (4) for the resulting amplitudes of these transformations. Let the $Z$-iterations be $\chi_0 \longrightarrow \chi_1 \longrightarrow\ldots\longrightarrow\chi_t$, $\chi_{i+1} = Z(\chi_i ) ,\ i=0,1,\ldots , t-1$; \begin{equation} \chi_i = b_i |e_1 e_2 \rangle +a_i |e_1 N_2 \rangle +\a_i |N_1 N_2 \rangle +\b_i |N_1 e_2 \rangle , \label{10} \end{equation} where $e_1$ and $e_1 ,e_2$ are the target states: unique solutions for $f_1 (x)=1$ and for $f_2 (x,y)=1$ respectively, $N_1 =\sum\limits_{i=2}^N e_i ,$ $N_2 =\sum\limits_{i\neq 2} e_i $ (we omit ancillary qubits). We represent the transformation $\chi_i \longrightarrow Z(\chi_i )$ as two sequential steps: $$ \chi_i \stackrel{1}{\longrightarrow} Z_1 (\chi_i )=\chi '_i \stackrel{2}{\longrightarrow} Z_2 (\chi '_i )=\chi_{i+1} , $$ where $Z_1 =-W_y R_{0y} W_y F_2 ,$ $Z_2 = -W_x R_{0x} W_x {\cal F}_1 .$ To calculate the change of amplitude resulting from the application of $Z_1$ (or $Z_2$) we shall fix a value of $x$ (or $y$ respectively). {\bf Step 1}. Denote amplitudes of basic states in $\chi '_i$ by the corresponding letters with primes: $$ \chi '_i = b'_i |e_1 e_2 \rangle +a'_i |e_1 N_2 \rangle +\a '_i |N_1 N_2 \rangle +\b '_i |N_1 e_2 \rangle . $$ Then for the two essentially different ways to fix a basic state for $x$: $x=e_1$ or $x+e_j ,\ j\neq 1$ we shall have the different expressions for new amplitudes. Use the property of the diffusion transformation $WR_0 W$ to be an inversion about average (look at \cite{Gr1}). Let $\lambda_{av}$ be an average amplitude of corresponding quantum state. a). \underline{$x=e_1$. } $\lambda_{av} = \frac{(N-1)a_i -b_i}{N} ,$ $b'_i =2\lambda_{av} +b_i$, $a'_i =2\lambda_{av} -a_i$, $$ \begin{array}{cl} b'_i =\frac{2(N-1)a_i -2b_i}{N} +b_i &=b_i(1-\frac{2}{N} )+2a_i (1- \frac{1}{N} ),\\ a'_i =\frac{2(N-1)a_i -2b_i}{N} -a_i &=-b_i \frac{2}{N} )+a_i (1- \frac{2}{N} ). \end{array} $$ b). \underline{$x=e_j ,\ j\neq 1$. } $\lambda_{av} =\frac{(N-1)\a_i +\b_i}{N} ,$ $\a '_i =2\lambda_{av} -\a_i$, $\b '_i =2\lambda_{av} -\b_i$, $$ \begin{array}{cl} \a '_i =\frac{2(N-1)\a_i +2b_i}{N} -a_i &=\a_i (1-\frac{2}{N} ) +2\b_i \frac{2}{N} ),\\ \b '_i =\frac{2(N-1)\a_i +2\b_i}{N} -\b_i &=\a_i (1-\frac{1}{N} )-\b_i (1-\frac{2}{N} ). \end{array} $$ {\bf Step 2}. $\chi '_i \stackrel{2}{\longrightarrow} Z_1 (\chi '_i )=\chi_{i+1} .$ We have two different ways to fix a basic state for $y:$ $y=e_2$ or $ y=e_j ,\ j\neq 2$. a). \underline{$y=e_2$. } $$ \lambda_{av} =\frac{(N-1)\b '_i -b'_i}{N}, \begin{array}{cl} b_{i+1} &= 2\lambda_{av} +b'_i =b'_i (1-\frac{2}{N} )+2\b '_i (1-\frac{1}{N} ),\\ \b_{i+1} &=2\lambda_{av} -\b '_i =\b '_i (1-\frac{2}{N} )-b'_i \frac{2}{N}. \end{array} $$ b). \underline{$y=e_j ,\ j\neq 2$.} $$ \lambda_{av} =\frac{(N-1)\a '_i -a'_i}{N}, \begin{array}{cc} a_{i+1} &= 2\lambda_{av} +a'_i =a'_i (1-\frac{2}{N} )+2\a '_i (1-\frac{1}{N} ),\\ \a_{i+1} &=2\lambda_{av} -\a '_i =\a '_i (1-\frac{2}{N} )-a'_i \frac{2}{N}. \end{array} $$ Hence, the recurrent formulas for amplitudes of sequential steps 1,2 acquire the following form: $$ \begin{array}{cl} b_{i+1} &= b_i (1-\frac{2}{N} )^2 +2a_i (1-\frac{1}{N} )(1-\frac{2}{N} )+ 4\a_i (1-\frac{1}{N} )^2 -2\b_i (1-\frac{2}{N} )(1-\frac{1}{N});\\ a_{i+1} &=a_i (1-\frac{2}{N} )^2 -b_i \frac{2}{N} (1-\frac{2}{N} )+2 \a_i (1-\frac{2}{N})(1-\frac{1}{N}) +2\b_i \frac{2}{N} (1-\frac{1}{N});\\ \a_{i+1} &=\a_i (1-\frac{2}{N})^2 +\b_i \frac{2}{N} (1-\frac{2}{N})-a_i (1-\frac{2}{N})\frac{2}{N} +b_i \frac{4}{N^2}; \\ \b_{i+1} &=2\a_i (1-\frac{1}{N})(1-\frac{2}{N}) -\b_i (1-\frac{2}{N})^2 -b_i (1-\frac{2}{N})\frac{2}{N} -2a_i (1-\frac{1}{N})\frac{2}{N} . \end{array} $$ Thus the matrix of one step of algorithm has the form $$ Z=\left( \begin{array}{cccc} 1 &2 &4 &-2\\ -\frac{2}{N} &1 &2 &\frac{4}{N}\\ \frac{4}{N^2} &-\frac{2}{N} &1 &\frac{2}{N} \\ -\frac{2}{N} &\frac{4}{N} &2 &-1 \end{array} \right) . $$ The system of recurrent equations can be rewritten as the following system of difference equations. \begin{equation} \begin{array}{cl} b_{i+1} -b_i &=2a_i +4\a_i -2\b_i +b_i O_1 (\frac{1}{N}) +a_i O_2 (\frac{1}{N})+\a_i O_3 (\frac{1}{N})+\b_i O_4 (\frac{1}{N});\\ a_{i+1} -a_i &= -\frac{2}{N} b_i +2\a_i +a_i O_5 (\frac{1}{N}) + b_i O_6 (\frac{1}{N^2}) +\a_i O_7 (\frac{1}{N}) +\b_i O_8 (\frac{1}{N});\\ \a_{i+1} -\a_i &=-\frac{2}{N} a_i +\b_i O_{13}(\frac{1}{N}) +a_i O_{14} (\frac{1}{N^2})+b_i O_{15}(\frac{1}{N^2})+\a_i O_{16}(\frac{1}{N}) ;\\ \b_{i+1} -\b_i &=-\frac{2}{N} b_i +2\a_i -2\b_i +a_i O_9 (\frac{1}{N}) + \a_i O_{10} (\frac{1}{N})+\b_i O_{11}(\frac{1}{N}) +b_i O_{12} (\frac{1}{N^2}) . \end{array} \label{11} \end{equation} \subsection{An approximation of amplitude's evolution by differential equations} Let $\{\bar c_i \}$ be a sequence of vectors from ${\rm C}^k :$ $\bar c_i =(c_i^1 ,c_i^2 ,\ldots ,c_i^k ),$ $c_i^j \in \rm C$, which satisfies the following system of difference equations \begin{equation} \bar c_{i+1} -\bar c_i =A\bar c_i, \label{12} \end{equation} where $A$ is a matrix of size $k\times k$ with complex elements. Let $m$ be an integer and a function $C(t):\ {\rm R}\longrightarrow{\rm C}^k$ is a solution of the system of differential equations \begin{equation} \dot C(t) =mAC(t) \label{13} \end{equation} with the initial condition \begin{equation} C(0)=\bar c_0 . \label{14} \end{equation} Then the exact solution of the Cauchy problem (13),(14) is $C(t)=R(t)\bar c_0 ,$ where the resolvent matrix $R(t) =\exp (mAt)$. The system (12) will be the system of difference equations approximating $C(t)$ by Euler's method if we consider $\bar c_i$ as an approximation of $C(i/m),\ i=0,1,\ldots .$ The accuracy of approximation may be obtained by the Taylor formula $C(\frac{i+1}{m})=C(\frac{i}{m})+\frac{1}{m} \dot C(\frac{i}{m})+ \frac{1}{2m^2}\ddot C(t_1 ),\ \frac{i}{m}<t_1 <\frac{i+1}{m}$. Here the error $\epsilon_1$ of the one step of the recursion (12) is the third summand $\frac{1}{2m^2} \ddot C(t_1 )=\frac{1}{2} A^2 C(t_1 )$. Thus the error at the first step is $\frac{1}{2} A^2 \exp (mA\theta_1 ) \bar c_0$, at the second step: $\frac{1}{2} A^2 \exp (mA\theta_2 ) \bar c_1 + \exp (mA\frac{1}{m}) \frac{1}{2} A^2 \exp (mA\theta_1 )\bar c0$ $= \frac{1}{2} A^2 \exp (mA\theta_2 ) (\bar c_0 +A\bar c_0 +\frac{1}{2} A^2 \exp (mA\theta_1 )\bar c_0 )+\exp (A) \frac{1}{2} A^2 \exp (mA\theta_1 ) \bar c_0$, etc., at the $i$-th step the error will be $\epsilon_i \leq\frac{3}{2}\sum\limits_{j=1}{i} \exp (A\a_j )A^2 \bar c_0$, where $0<\a_j <1$. Hence if $\| \bar c_0 \| \leq h$, then the error after $i$-th step is $\epsilon_i =O(ih)$. Particularly, for the initial conditions $\|c_0 \| =O(\frac{1}{N} )$ a good approximation can be obtained if $i=o(N)$, and thus we can solve the Cauchy problem instead of (11) for $i=O(\sqrt{N} )$ having error as small as required for sufficiently large $N$. Define a new function $B(\tau )$ as follows $B (t m )=C(t)$. In terms of $B$ the Cauchi problem (13), (14) acquires the form \begin{equation} \frac{{\rm d}}{{\rm d}\tau} B(\tau)=AB(\tau ),\ B(0)=c_0 . \label{15} \end{equation} Apply this to the solution $\bar c_i =|b_i ,a_i ,\a_i ,\b_i \rangle$ of (11), where $\bar c_0 = |\frac{1}{N} ,\frac{1}{N} ,\frac{1}{N} ,\frac{1}{N}\rangle$. Put $B=|b,a,\a ,\b \rangle$ for the scalar functions $b,a,\b ,\a$ and denote the argument of the function $B$ by $t$. Then the equation (15) approximating (11) acquires the following form: \begin{equation} \begin{array}{cl} \dot b &=2a+4\a -2\b +bO_1 (\frac{1}{N} )+\epsilon_1 +aO_0 (\frac{1}{N});\\ \dot a &=-\frac{2}{N} b+2\a +\epsilon_2 +O_2 (\frac{1}{N})a;\\ \dot \b &=-\frac{2}{N} b+2\a -2\b +\epsilon_4 +O_4 (\frac{1}{N}) a;\\ \dot \a &=-\frac{2}{N} a +\epsilon_3, \end{array} \label{16} \end{equation} where $\epsilon_i =a O_{0i} (\frac{1}{N^2}) +bO_{1i} (\frac{1}{N^2}) +\b O_{2i}(\frac{1}{N}) +\a O_{3i}(\frac{1}{N}),\ i=1,2,3,4,$ with the initial condition \begin{equation} b(0)=a(0)=\b (0)=\a (0)=\frac{1}{N} . \label{17} \end{equation} Then for $t=O(\sqrt{N}),\ i=[t]$ the vector of error will be $\bar\delta =\bar B(t) -\bar c_i =O(1/\sqrt{N}),\ \ N\longrightarrow\infty$ and with this accuracy we can write $b(i)\approx b_i$ for the amplitude $b_i$ of target state $|e_1 ,e_2 \rangle$. \subsection{Tight analysis of the parallel quantum algorithm for RS} Now we shall take up the system of linear differential equations (16) with the initial conditions (17). Our goal is to solve it on a segment of the form $0\leq t\leq O(\sqrt{N}).$ The system (16) can be represented in the form $\dot B =MB$, where its matrix $M=Z-1=\tilde A_0 +E+H$ ($1$ denotes the identity matrix) for the matrices $$ \tilde A_0 =\left( \begin{array}{cccc} 0 &2 &4 &0\\ -\frac{2}{N} &0 &2 &0\\ 0 &-\frac{2}{N} &0 &0 \\ 0 &0 &0 &0 \end{array} \right), E=\left( \begin{array}{cccc} 0 &0 &0 &0\\ 0 &0 &0 &0\\ 0 &0 &0 &0 \\ -\frac{2}{N} &0 &2 &-2 \end{array} \right), H=\left( \begin{array}{cccc} d_1 &d_1 &d_1 &-2+d_1\\ d_2 &d_1 &d_1 &d_1\\ d_2 &d_2 &d_1 &d_1 \\ d_2 &d_1 &d_1 &d_1 \end{array} \right), $$ where $d_l$ denotes different expressions of the form $O(N^{-l} ),\ l=1,2.$ We shall show that the deposit of $\tilde A_0$ to the solution of (16),(17) is the main and deposits of $E$ and $H$ are negligible. What is the main difficulty here? Consider the resolvent matrix for the Cauchy problem (16), (17), this is the solution $R(t)$ of the differential equation for matrices: $\dot R=MR$ with the initial condition $R(0)=1$. Then we have $C(t)=RC(0)$. The matrix $R$ has the form $\exp (Mt)$. But in our case the matrices $\tilde A_0 ,E,H$ do not commutate, hence we cannot use the standard properties of exponent. In order to cope with this task we shall at first solve the Cauchy problem at hand neglecting deposits of $E$ and $H$ to the main matrix $M$. The legality of this approximation is shown in the Appendix. Now take up the reduced equation $\dot C(t) = AC(t)$ with the initial condition $C(0)=c_0$. Excluding the last column and row containing only zeroes we obtain the new matrix $A_0$. The characteristic equation for $A_0$ is $\lambda^3 +\frac{8}{N}\lambda -\frac{16}{N^2}=0$ and its nonzero solutions in within $O(\frac{1}{N})$ are $\lambda_{1,2} =\stackrel{+}{-} 2\sqrt{2}i/\sqrt{N}$. Then standard calculations give the approximation of solution as \begin{equation} \begin{array}{cl} b &=\frac{1}{2} -\frac{1}{2}\cos \frac{2\sqrt{2}t}{\sqrt{N}},\\ a &=\frac{1}{\sqrt{2N}}\sin \frac{2\sqrt{2}t}{\sqrt{N}}+\frac{1}{N} \cos\frac{2\sqrt{2}t}{\sqrt{N}},\\ \a &=\frac{1}{2N} \cos\frac{2t}{\sqrt{N}}+\frac{1}{2N} \end{array} \label{18} \end{equation} in within $|O(\frac{1}{\sqrt{N}}), O(\frac{1}{N} ), O(\frac{1}{N\sqrt{N}}) \rangle$. The amplitude $b$ from (18) peaks in the point $t_1 =\frac{\pi\sqrt{N}}{2\sqrt{2}}$ where $b(t_1 )=1$ in within $O(\frac{1}{\sqrt{N}})$. Assuming that deposits of $E$ and $H$ to the solution are small, we obtain that the amplitude of target state $|e_1 ,e_2 \rangle$ will be $1-O(1/\sqrt{N} )$ after $\left[\frac{\pi\sqrt{N}}{2\sqrt{2}} \right]$ steps of parallel algorithm which is $\sqrt{2}$ times as small as the time of sequential quantum search. \subsection{Completion of the proof} In the first version of this paper the deposit of $E$ and $H$ was estimated by the conventional procedure of approximation a solution of differential equation (look at the Appendix). This is the immediate but cumbersome way to prove that this deposit is vanishing. After the publication of the first version of this paper Farhi and Gutmann informed me how to simplify this construction by uniting by pairs the sequential transformations in parallel algorithm (\cite{FG1}). In this section I combine this idea with the approach of the first version. At first turn to the orthonormal basis $E_1 = |e_1 e_2 \rangle$, $E_2 =\frac{1}{\sqrt{N}}|e_1 N_2 \rangle$, $E_3 =\frac{1}{N}|N_1 N_2 \rangle$, $E_4 =\frac{1}{\sqrt{N}}|N_1 e_2 \rangle$. The matrix $Z$ in this basis acquires the form $$ A_1 =\left( \begin{array}{cccc} 1 &\frac{2}{\sqrt{N}} &\frac{4}{N} &-\frac{2}{\sqrt{N}}\\ -\frac{2}{\sqrt{N}} &1 &\frac{2}{\sqrt{N}} &\frac{4}{N}\\ \frac{4}{N} &-\frac{2}{\sqrt{N}} &1 &\frac{2}{\sqrt{N}} \\ -\frac{2}{\sqrt{N}} &\frac{4}{N} &\frac{2}{\sqrt{N}} &-1 \end{array} \right) . $$ Now group together each pair of unitary transformations in the algorithm: $\chi_{2k} \longrightarrow\chi_{2k+1} \longrightarrow\chi_{2k+2}$ and denote by $B$ the corresponding matrix: $B_0 =A_1^2$. It is sufficient to prove that $\| B_0^{[\frac{\pi\sqrt{N}}{4\sqrt{2}}]}|0,0,1,0\rangle -|1,0,0,0\rangle\| =O(\frac{1}{\sqrt{N}})$, because one application of $A_1$ can only increase the error by $O(\frac{1}{\sqrt{N}})$. The Cauchy problem for the recursion $\bar c_{i+1} = B_0 \bar c_i$ has the form $\dot{\bar c} =(B_0 -1)\bar c,$ $\bar c=\bar c_0$, and its resolvent has the form $R=\exp Bt$, where $B=B_0 -1$. We have: $$ B \approx\frac{4}{\sqrt{N}} \left( \begin{array}{cccc} 0 &1 &0 &0 \\ -1 &0 &1 &0 \\ 0 &-1 &0 &0 \\ 0 &0 &0 &0 \end{array} \right) $$ in within $O(\frac{1}{N})$. Thus we can consider only a projection of $\bar c$ to the subspace ${\cal H}_1$ spanned by $E_1 ,E_2 ,E_3$. Denote by $D$ the matrix $$ \left( \begin{array}{ccc} 0 &-\frac{i}{\sqrt{2}} &0\\ \frac{i}{\sqrt{2}} &0 &-\frac{i}{\sqrt{2}}\\ 0 &\frac{i}{\sqrt{2}} &0 \end{array} \right) . $$ Then the restriction of $B$ to ${\cal H}_1$ has the form $\frac{4\sqrt{2}i}{\sqrt{N}}D$. It is easily seen that $D^{2k+1}=D$, $D^{2k} =D^2$ for $k=1,2,\ldots$. If the number of steps is $\left[\frac{\pi\sqrt{N}}{2\sqrt{2}}\right]$ then $t=\left[\frac{\pi\sqrt{N}}{4\sqrt{2}}\right]$. Here in within $O(\frac{1}{\sqrt{N}})$ we have $$ \begin{array}{cc} |b,a,\a\rangle &\approx \exp(\frac{4\sqrt{2}i}{\sqrt{N}} Dt)= \exp (\pi i D)= \cos (\pi D)+i\sin(\pi D)\\ &=1-\frac{(\pi D)^2}{2}+\frac{(\pi D)^4}{4!}-\ldots +i(\pi D- \frac{(\pi D)^3}{3!}+\ldots )\\ &=1-D^2 (1-\cos\pi )+iD\sin\pi =1-2D^2 \end{array} $$ that is $$ \left( \begin{array}{ccc} 0 & 0 & 1\\ 0 & -1 & 0\\ 1 & 0 & 0 \end{array} \right) . $$ The initial state is $|0,0,1\rangle$ in within $O(\frac{1}{\sqrt{N}})$. Application of this matrix to the initial state gives $|1,0,0\rangle$ with this accuracy. Theorem is proved. \section{Parallel implementation of iterated quantum search} \subsection{Parallel quantum algorithm for IS} Now take up an IS problem for arbitrary $k$. Consider an evolution of amplitudes arising when $k$ oracles work in parallel. Let $\chi_i = a_0^i |N_1 ,N_2 ,\ldots ,N_k \rangle + a_1^i |e_1 , N_2 ,\ldots ,N_k \rangle +\ldots +a_k^i | e_1 ,e_2 ,\ldots , e_k \rangle +R_i$ (it generalizes (10)), where $R_i$ contains only basic states of the form \newline $|\ldots ,N_p , \ldots ,e_q ,\ldots \rangle$. The natural generalization of the transformation (8) will be $$ Z_k =(-1)^k {\cal W}^{(k)} {\cal R}_0^{(k)}{\cal W}^{(k)} {\cal F}^{(k)}, $$ where ${\cal W}^{(k)} = W_1 \bigotimes W_2 \bigotimes \ldots \bigotimes W_k \bigotimes I$, each W-H transformation $W_i$ acts on $x_i$, $i=1,2,$ $\ldots ,$ $ k$, ${\cal R}_0^{(k)}= R_{01} \bigotimes \ldots \bigotimes R_{0k} \bigotimes I$, each rotation of 0's phase $R_{0i}$ acts on $x_i$, $i=1,2,\ldots ,k$, ${\cal F}^{(k)}= F_1 \bigotimes\ldots \bigotimes F_k \bigotimes I$, each $F_i$ acts on $x_i$ and inverses the sign of $e_i$, identities $I$ act on ancilla. Let a matrix $A$ determines an evolution of quantum state in parallel algorithm such that $\chi_i =A \chi_{i-1}$. $A$ represents the operator in $2^{kn}$-dimensional space. We reduce $A$ to an operator $A_r$ acting on $k+1$-dimensional space generated by the vectors \newline $|N_1 ,N_2 , \ldots ,N_k \rangle , |e_1 , N_2 ,\ldots ,N_k \rangle ,\ldots , | e_1 , e_2 ,\ldots , e_k \rangle$. Then represent $A_r$ as $A_r =A_0 +B$, where $A_0$ is Jacobi matrix of the form \begin{equation} \left( \begin{array}{ccccc} 0 &2 &0 &\ldots &0\\ -\frac{2}{N} &0 &2 &\ldots &0\\ \ldots &\ldots &\ldots &\ddots &\ldots\\ 0 &\ldots &-\frac{2}{N} &0 &2\\ 0 &\ldots &0 &-\frac{2}{N} &0\\ \end{array} \right) . \label{27} \end{equation} This matrix has the following nonzero elements: 2 - above the main diagonal and $-\frac{2}{N}$ - behind it. Its size is $(k+1)\times (k+1)$. Assume that the effect of reduction: $A$ to $A_r$ and the deposit of $B$ are negligible. Generally speaking for $k>2$ this assumption should be proved. An evolution of amplitude can be represented approximately as the solution of Cauchy problem \begin{equation} \dot{\bar a} = A\bar a ,\ \ a(0) =|N^{-k/2} ,\ldots ,N^{-k/2} \rangle , \label{28} \end{equation} where $\bar a (i) \approx |a_1^i ,\ldots , a_k^i \rangle$, $i$ integer, we assume that $k\ll \sqrt{N}$. Given eigenvalues of matrix (19), a general solution can be obtained by the standard procedure. These eigenvalues for Jacobi matrix is known (look at \cite{Bel}, Chapter 2, ex. 32), there are $\lambda_m =-\frac{4i}{\sqrt{N}} \cos m\theta ,\ \ m=1,2,\ldots ,k, \theta =\frac{\pi}{k+2}$. We shall not solve (20) here. \subsection{Perspectives of parallel quantum algorithm for iterated search} One can ask: can we obtain a speedup by a big constant factor for iterated quantum search when applying parallel action of more than two oracles? In all probability the answer is no. Estimate the growth of target amplitude from above. Canceling all $-\frac{2}{N}$ we can only increase this growth. This results in a simple system of linear differential equations whose solution is $a_k (t)=N^{-k/2} +2^k {\int\limits_0^t}^{\{ k\}} N^{-k/2} dt = N^{-k/2} + 2^k N^{-k/2} \frac{t^k}{k!}$. The parallel algorithm for IQS can exceed sequential quantum search only if $a_k (t)$ is substantially large for $t < k\sqrt{N}$. For example let the parallel algorithm work only the quarter of time required for the sequential search. Then it can not reach the vanishing error probability for any $k$, because for such $t=\frac{\pi\sqrt{N}k}{16}$ $a_k \approx \left(\frac{\pi}{8}\right)^k \frac{k^k}{k!} <1$. Nevertheless, parallel quantum algorithm has one advantage. Compare sequential and parallel quantum algorithms for IS in case $1\ll k\ll\sqrt{N}$, if total time $t=\sqrt{N}$. If $t=\sqrt{N}$ then $t$ sequential applications of $Z^{(k)}$ raise the amplitude $a_k$ up to the value $a_k (t)=\frac{a^k}{k!} ,\ \ a=2-\epsilon$, where $\epsilon\longrightarrow 0 \ \ (N\longrightarrow\infty)$. Hence the resulting probability is $P_{par} = \left(\frac{a^k}{k!}\right)^2$. On the other hand if we have a total time $\sqrt{N}$ then we can apply sequential quantum searches with the time $\sqrt{N}/k$ for each $x_i^0 , \ \ i=1,2,\ldots ,k$. The probability $P_{seq}$ to find $x_k^0$ will be less than $\left(\frac{2}{k}\right)^{2k}$ because for one search it does not exceed $\left(\frac{2}{k}\right)^2$. Consequently, $P_{par}$ exceeds $P_{seq}$ in more than $2^{2k}$ times. \subsection{Conclusion} To sum up, the parallel algorithm constructed above for repeated quantum search is $\sqrt{2}$ times as fast as sequential application of fast quantum search and it requires the same hardware. The advance is taken of interference arising when two oracles act simultaneously on the set of entangled qubits. This parallel quantum algorithm can be applied to the problem of $k$ dependent iterations of quantum search in areas of $N$ elements each, with the same effect of speedup in $\sqrt{2}$ times. Here the error probability will be vanishing if $k=o(\sqrt{N}),\ N\longrightarrow\infty$. In addition, for the fixed total time $\sqrt{N}$ a probability of success for parallel algorithm is $2^{2k}$ times as big as for sequential algorithm. The effect of speedup in $\sqrt{2}$ times can not be increased essentially by the same procedure if we increase a number of oracles involved in the simultaneous action. Nevertheless, a possibility of further speedup of the iterated quantum search still remains. \section{Acknowledgements} I am grateful to Victor Maslov for his support, to Oleg Khrustalev, Anatolyi Fomenko, Voislav Golo, Grigori Litvinov, Yuri Solomentsev and Kamil Valiev for the interesting discussions and help. I am especially grateful to Lov Grover for his attention to my work and useful discussions and to Peter Hoyer for valuable comments to my previous works. I also thank Charles Bennett, David DiVincenzo, Lev Levitin, Richard Cleve, Richard Jozsa and others for the fruitful discussions at the conference NASA QCQC'98. I would like also to express my gratitude to Edward Farhi and Sam Gutmann who have read the first version of this work and suggested the simplification of the proof.
\section{Introduction} Molecular outflows associated with young stellar objects are mostly made of ambient material entrained by a primary wind from the central source. While young molecular outflows are highly collimated with HV jets (see e.g. Bachiller, 1996), more evolved outflows are poorly collimated with shell-like structures (see e.g. Snell et al., 1980; Fuente et al., 1998). Different kinds of models (wind-driven bubbles and steady state jets) have been developed to explain the two types of outflows (see e.g. Cabrit, 1995) but none of them can account for all the observational properties. A jet variable in velocity and/or direction, would explain the momentum distribution (Chernin \& Masson, 1995), the multiple acceleration sites (see e.g. Bachiller, 1996), and the evidences of a wiggling molecular outflow (Davis et al., 1997). Furthermore, models of the interaction of a jet variable in time and direction predict that the jet breaks, given rise to independent HV bullets located in a shell-like structure with a non-negligible transverse velocity component (see e.g. Raga \& Biro, 1993). Thus, observations of molecular outflows oriented along the line of sight and powered by a variable jet should show a ring-like structure of HV jet-bullets, and would allow to test these kind of models. In this letter, we present high sensitivity CO observations of the molecular outflows in the Orion A molecular cloud (IRc2, see e.g. Wilson et al.,1986; and Orion-S, Schmid$-$Burgk et al., 1990). The morphology of the HV CO emission around the IRc2 outflow reveals the presence of a bipolar structure $20''$ from I source surrounded by a HV, ring-like structure of CO bullets that cannot be explained by the weakly collimated bipolar outflow model proposed by Chernin \& Wright (1996), and suggest a jet driven molecular outflow oriented along the line of sight (Johnston et al., 1992). The combination of our results with those of the H$_2$O masers, strongly supports the idea that these molecular outflows are driven by jets which change in direction with time. \section{Observations} The observations of the $J=2\rightarrow1$ lines of CO where carried out with the IRAM 30-m telescope at Pico Veleta (Spain). The observation were made with a SIS receiver tuned to single side band (SSB) with an image rejection of $\sim 8\,$dB. The SSB noise temperatures of the receivers at the rest frequency was 300$\,$K. The half power beam width of the telescope was $12''$. For spectrometers, we used a filter banks of $512\times1\,$MHz that provided a velocity resolution of 1.3$\,$km$\,$s$^{-1}$. The observation procedure was position switching with the reference taken at a fixed position located 15$'$ away in right ascension. The mapping was carried out by combining 5 on-source spectra with one reference spectrum. The typical integration times were 20$\,$sec for the on-positions and 45$\,$sec for the reference spectra. Pointing was checked frequently on nearby continuum sources and Jupiter, and the pointing errors were $\leq4''$. The calibration of the data was made by observing the sky, a hot and a cold load. The line intensity scale has been converted to units of main beam brightness temperatures by using a main beam efficiency of 0.45. The RMS noise of the map was 0.6$\,$K. \section{HV bullets in the IRc2 and Orion-S outflows} \subsection{Spectral features} \begin{figure*} \vspace{8.5cm} \special{psfile=ori-lan-f1.ps hoffset=0 voffset=-207 hscale=75 vscale=75 angle=0} \caption{Left panels: a) sample CO $J=2\rightarrow1$ line profiles taken towards selected positions in the vicinity of the molecular outflow surrounding IRc2. The offsets shown in the upper right corner of each box are relative to IRc2. The vertical arrows show the location of HV ``bullets'' similar to those observed in some bipolar outflows driven by low mass stars. Central panels: b) and c) integrated intensity of the CO $J=2\rightarrow1$ line emission between $-90$ and $-10\,$km$\,$s$^{-1}$, and between $40$ and $90\,$km$\,$s$^{-1}$ for the blue and red HV bullets respectively. The maps have been obtained by subtracting a Gaussian profile to the broad line wings. The first contours level is $2\,$K$\,$km$\,$s$^{-1}$, and the interval between levels is $7\,$K$\,$km$\,$s$^{-1}$. d) and e) integrated intensity maps of the CO $J=2\rightarrow1$ line over the most extreme velocities (``molecular jet''), from $-110$ to $-90\,$km$\,$s$^{-1}$ for the blue jet, and from $95$ to $115\,$km$\,$s$^{-1}$ for the red jet. For these two panels, the first contour level is $2\,$K$\,$km$\,$s$^{-1}$ and the interval between levels is $1\,$K$\,$km$\,$s$^{-1}$. The circle in the lower left panel represents the size of the beam. For all the panels, the filled star represents the position of IRc2, and triangles and dots represent, respectively, the positions of the high and low velocity H$_2$O maser taken from Gaume et al. (1998), and the filled squares the positions of some H$_2^*$ features (Stolovy et al., 1998).} \label{fig:ring} \end{figure*} The left panel of Fig. \ref{fig:ring} (panel a) shows a sample of spectra taken toward the IRc2 outflow. The profiles show the typical broad line wings ($\pm 100\,$km$\,$s$^{-1}$) associated with this molecular outflow. Superposed on these, because of the better sensitivity than previous published data, we have detected well defined spectral HV features restricted to certain radial velocity ranges (see the vertical arrows in the spectra of Fig. \ref{fig:ring}a). Most of the HV features in the IRc2 outflow appear at radial velocities between 30 and $90\,$km$\,$s$^{-1}$. Similar HV features are also clearly identified in the spectra of Orion-S outflow (left panel of Fig. \ref{fig:ori-s}; Schmid-Burgk, private communication ). The HV spectral features detected in both molecular outflows are reminiscent of the HV bullets found in the molecular outflows driven by low mass stars (see Bachiller, 1996). Furthermore, the CO profiles towards the IRc2 outflow are similar to the recently discovered H$_2^*$ bullets (Stolovy et al., 1998). The CO HV features reported here represents the first detection of HV bullets in molecular outflow powered by a massive star. \subsection{Morphology} The upper right panels (b and c) of Fig. \ref{fig:ring}, and the right panel of Fig. \ref{fig:ori-s} show, respectively, the spatial distribution of the integrated intensity of the blue and redshifted CO HV bullets in the IRc2 and Orion-S outflows. The spatial distribution of the HV bullets have been obtained by subtracting the smooth broad velocity component by fitting a Gaussian profile. It is remarkable that the blue and redshifted CO HV bullets around IRc2 (Fig. \ref{fig:ring}b, and c) are distributed in an elliptical ring-like structure with a size of $\sim 10''\times50''$ (0.02$\times0.1\,$pc at the distance of 0.5$\,$Kpc) with IRc2 located in the southeast edge of the HV bullets rings. The ring morphology of the CO HV bullets in the IRc2 outflow shows only minor changes with the radial velocity, indicating a nearly uniform distribution over the whole velocity range. Although the typical thickness of the rings is $12''-20''$ ($0.02-0.04\,$pc), some positions are unresolved (thicknesses $\leq6''$; 0.01$\,$pc). This suggests that the blue and redshifted CO HV bullets are generated in a thin layer of HV gas distributed in a ring-like structure. It is interesting to note that the HV bullet rings are broken in the northwest edge just at the bottom of the H$^*_2$ fingers (Allen \& Burton, 1993). This indicates that the H$^*_2$ fingers might have been produced when the hot gas within the shocked region breaks into the more diffuse medium and rapidly expands (McCaughrean \& Mac Low, 1997). In panels d and e of Fig. \ref{fig:ring} we also show the spatial distribution of the CO emission for the most EHV components ($\geq\mid90\mid\,$km$\,$s$^{-1}$). The bulk of the blueshifted and redshifted EHV gas is located $20''$ north of source I. None of the current jet and wind models can account of all the observational properties of the IRc2 outflow. The morphology of the CO emission at moderate velocities favors a biconical outflow structure that has a wide ($130^{\rm o}$) opening angle (Chernin \& Wright, 1996) powered by source I (Menten \& Reid, 1995). The morphology of the SiO maser spots near source I is consistent with a wide angle biconical outflow, but this simple model cannot account for the H$_2$O maser emission (Greenhill et al., 1998; Doeleman et al., 1999). The morphology of the HV CO bullet ring-like structures roughly trace the edges of the proposed biconical structures. However, the bipolar distribution of the EHV gas emission $20''$ north of source I, surrounded by the CO HV bullet rings is inconsistent with a wide angle biconical outflow model. In the next section, we analyze the alternative model of a jet driven molecular outflow directed along the line of sight (Johnston et al., 1992). Fig. \ref{fig:ori-s}b shows the spatial distribution of the HV bullets for the Orion--S outflow. The HV bullets are small condensations (size $\sim30''$; 0.07$\,$pc) and show the typical bipolar distribution with the blue and redshifted features spatially separated from the powering source. The HV bullets basically outlines the full extent of this outflow. The outflow containing the CO bullets reported in this letter is perpendicular to the low velocity redshifted outflow found by Schmid-Burgk et al. (1990), and consistent with the spatial distribution of the SiO outflow found by Ziurys et al. (1990). The possible driving source, as defined by the center of symmetry from the kinematics of the HV gas, must lie at a position $\sim18''$ north from FIR$\,4$ where no prominent continuum source has been detected so far (Mundy et al., 1986; Wilson et al., 1986). In this outflow, the blue HV bullet shows different spatial distribution and velocity than the redshifted one. While the red CO bullet has a moderate radial velocities of $\sim60\,$km$\,$s$^{-1}$ and is close to the exciting source, the blue CO bullet has very high velocity ($\sim100\,$km$\,$s$^{-1}$) and is located further away from the exciting source. The different distribution might be due to the fact that the blue bullet is less massive than the red one, and it has been already accelerated up to the terminal velocity. In fact, the Orion--S outflow is rather young with a dynamical age of only $10^3\,$years. \section{Discussion} \subsection{Jet driven molecular outflow in Orion A} With the present data we conclude that the CO HV bullet rings around IRc2 represent thin layers of HV condensations which have been shocked and accelerated by a fast jet oriented along the line of sight. The orientation of the flow along the line of sight is also suggested by the kinematics of the SiO masers around source I (Doeleman et al., 1999). In addition to the morphological arguments, the location of different shock tracers in this region like the H$_2$O masers, their kinematics, and the H$_2^*$ features can be accounted by this model. Fig. \ref{fig:ring}b, c, d and e, show the location of the low (filled circles), the high velocity (open triangles) H$_2$O masers, and the H$_2^*$ bullets (filled squares) adapted from Stolovy et al. (1998). The shock tracers, the low velocity H$_2$O masers and the H$_2^*$ bullets, are located in the inner border of the CO HV bullets ring and, therefore, are surrounding the EHV jet. The H$_2^*$ bullets and the H$_2$O masers are displaced from the CO HV bullets by $\sim10''$ ($2\times10^{-2}\,$pc). On the other hand, the high velocity H$_2$O masers are well correlated with the EHV jet indicating that they, indeed, arise from the interaction of the jet with gas which was already accelerated close to the terminal velocity of the outflow. \begin{figure} \vspace{4.3cm} \special{psfile=ori-lan-f2.ps hoffset=-182 voffset=-178 hscale=75 vscale=75 angle=0} \caption{Left panels: a) sample of CO $J=2\rightarrow1$ spectra taken towards selected positions of the molecular outflow Orion--S. The offsets shown in the upper right corner of each box are relative to the position of FIR$\,$4. The vertical arrows show the location of HV ``bullets''. Right panel: b) integrated intensity maps of the CO $J=2\rightarrow1$ line in the Orion--S outflow. The offsets are relative to the position of FIR$\,$4. The intervals of velocity integration are: from $-150$ to $-5\,$km$\,$s$^{-1}$ for the blue wing, and from 25 to 100$\,$km$\,$s$^{-1}$ for the red wing. For both wings the first contour level is $9\,$K$\,$km$\,$s$^{-1}$ and the interval between levels is $2.5\,$K$\,$km$\,$s$^{-1}$. The circle in the lower left panel represents the size of the beam. The filled triangles and dots represent, respectively, the positions of the blue and red H$_2$O maser taken from Gaume et al. (1998). The filled square shows the position of the possible exciting source.} \label{fig:ori-s} \end{figure} The observed morphologies of the CO bullet ring, the shock tracers, and the EHV gas can be explained by a fast jet moving along the line of sight and interacting with the surrounding molecular gas. In this scenario, the fast jet with material moving at velocities $\geq100\,$km$\,$s$^{-1}$ will interact with the surrounding ambient gas generating strong shocks that will compress, heat the gas, and even photodissociate molecules in its surroundings and in the head of the jet. As one moves away from the working surfaces of the jet, the material will cool down. First, H$_2$ molecules will be formed producing the strong H$_2^*$ features in the densest hot clumps, and also the H$_2$O maser emission. As the accelerated post-shock material moves further away from the interface region it will cool further forming CO molecules (see Hollenbach \& McKee, 1989) and given rise to the CO HV bullets. The stratification of the different shock tracers indicates that the shocked gas is not only accelerated along the line of sight, but also perpendicularly to the jet axis. This is consistent with the measured proper motion of the low velocity H$_2$O masers which indicates that these are expanding at a velocity of $18\,$km$\,$s$^{-1}$ (Genzel et al., 1981). This allows us to measure for the first time the transverse velocity in a jet driven molecular outflow which is $\sim20\%$ of the jet velocity. For a transverse velocity of $18\,$km$\,$s$^{-1}$, the separation between the CO bullet ring and the H$_2^*$ indicates that the CO bullets ring has gone through the shock $\sim10^3\,$years ago, which is at least one order of magnitude larger than the typical time required to cool down the material and to produce CO molecules efficiently (Hollenbach \& McKee, 1989). This indicates that the H$_2^*$ and the H$_2$O emissions trace very recent shocks produced less than 100$\,$years ago, and the CO HV bullets must have been produced by shocks more than $10^3\,$years ago. In contrast to the IRc2 outflow, the Orion--S outflow seems to be oriented nearly perpendicular to the line of sight. The combination of these two outflows powered by massive stars with strong H$_2$O maser emission, but oriented with very different angles to the line of sight allows a three dimensional study of the interaction of the HV jets with the ambient cloud. We now study the history of the young Orion--S outflow using the picture of the time dependent shock tracers obtained from the IRc2 outflow. In the case of the Orion--S outflow, all the H$_2$O masers are associated with the HV gas seen in CO (lower panel of Fig. \ref{fig:ori-s}b). The redshifted H$_2$O masers are on the western part of the red bullet facing the exciting source where a jet, nearly perpendicular to the line of sight, impinges. The interaction of the jet with the CO HV bullets is not only supported by the morphology, but also by the H$_2$O masers which have radial velocities similar to that of the red CO bullet. This indicates that the most recent shocks are indeed produced in the HV bullet. The blueshifted bullet seems to be in a different stage of evolution. As previously mentioned, the blue bullet is further from the exciting source than the red one, and the blueshifted masers are only found close to the exciting source at radial velocities close to the ambient velocities. This indicates that the most recent interaction of the jet is not occurring with the HV bullet material, but in ambient material which has not been yet affected by the jet, suggesting that the jet might have slightly changed the direction at which it is ejected. Therefore this can be understood in a model in which the molecular outflows are powered by a jet whose direction is changing with time (see e.g. Raga \& Biro, 1993). \subsection{The Origin of shell-like molecular outflows} The large amount of the molecular gas mass observed in outflows suggests that these are mostly made by ambient material entrained by a ``primary jet'' (Bachiller, 1996). Entrainment can be divided in two categories: prompt entrainment at the jet heads (bowshock), and steady-state entrainment along the side of the jet due to Kelvin-Helmholtz (KH) instabilities. Most of the studies of the lines profiles seem to indicate that the prompt entrainment at the jet heads is the main mechanism for molecular entrainment (Chernin et al., 1994). However, to explain the data, Chernin et al. (1994) proposed a jet/bowshock model in which the jet is variable in velocity and/or direction. In the proposed scenario of a jet along the line of sight powering the IRc2 outflow, changes in the direction of the jet would explain the ring-like structure of the CO bullets (Raga \& Biro, 1993), the spatial distribution of all the H$_2$O masers, as well as the large number of the H$_2$O masers at relatively low radial velocities. One important aspect of our interpretation is that the entrained material is moving perpendicular to the jet with velocities of up to $20\%$ of the jet velocity. This indicates that in a time scale of $10^5\,$years, the IRc2 molecular outflow will form a cavity around the driving source with typical size of $\sim3\,$pc. This is similar to those found in molecular outflows powered by intermediate mass stars (NGC$\,$7023; Fuente et al., 1998) and by low mass stars (L$\,$1551-IRS5; Snell et al., 1980). In summary, the morphology and the radial velocities of the CO HV bullets, the H$_2$O masers and the H$_2^*$ bullets in the Orion A outflows can be explained by the interaction of jet-driven molecular outflows powered by variable jets in direction with the ambient gas. \acknowledgements We thank Dr. A. Fuente for critical reading of the manuscript, and the referee, Dr. D.S. Shepherd, for her suggestions. Part of this work was supported by the Spanish DGES under grant number PB96-0104.
\section{Introduction} Transport properties of quantum dots at zero frequency have been extensively studied and by now many aspects are well understood \cite{Coulomb}. In some studies finite, but still low frequency signals were applied to a gate electrode nearby the quantum dot. Capacitance spectroscopy on quantum dots has been performed at kHz frequencies \cite{Ashoori}. At MHz frequencies, quantum dots can be operated as turnstiles or pumps \cite{Leopump}. These frequencies, $f$, are low in the sense that the photon energy, $hf$, is much smaller than the thermal energy, $k_{B}T,$ and thus the discrete photon character cannot be discerned. For sufficiently high frequencies, such that $ hf\gg k_{B}T$, the interaction of the electromagnetic field with the electrons confined in a quantum dot would be analogous to light spectroscopy studies on atoms. However, different quantum dots are not microscopically equal. The response of an ensemble of quantum dots to light excitation is therefore strongly averaged. Despite this averaging, excitation studies on arrays of quantum dots by far-infrared light (i$.$e$.$ the THz regime) have revealed the spectrum of collective modes \cite{Merkt}, i$.$e$.$ the sloshing modes of the whole electron puddle in the external potential. Excitations within the electron puddle are difficult to create since, according to the generalized Kohn theorem \cite{Merkt}, the dipole field of far-infrared radiation does not couple to the relative coordinates of electrons confined in a parabolic quantum dot. This problem is circumvented by inelastic light scattering experiments which have been able to detect electronic excitations in arrays of quantum dots that can be related to a discrete single-particle spectrum \cite{Raman}. Recently, this technique has also probed excitons in a single quantum dot \cite{Gerhard}. In this review, we discuss electron transport experiments on single quantum dots, that are irradiated by a microwave signal. In contrast to light transmission, luminescence, or inelastic light scattering measurements (see for a review Ref$.$ \cite{Hawrylak-book}), we measure the dc current in response to a microwave signal. Current can flow through a quantum dot when a discrete energy state is aligned to the Fermi energies of the leads. This resonant current is carried by elastic tunneling of electrons between the leads and the dot. An additional time-varying potential $\widetilde{V}\cos (2\pi ft)$ can induce {\it inelastic} tunnel events when electrons exchange photons of energy $hf$ with the oscillating field. This inelastic tunneling with discrete energy exchange is known as photon assisted tunneling (PAT). Microwave studies have a long tradition in the field of superconductor-insulator-superconductor tunnel junctions \cite{Tucker}, for which the theory was first described by Tien and Gordon in 1963 \cite{Tien}. Despite many proposals and a long search \cite{failures}, it took thirty years before PAT was also observed in a non-superconducting system. In 1993 PAT features were seen in the current-voltage characteristics of a GaAs/AlGaAs superlattice under THz irradiation from a free-electron laser \cite{SantaBarbara}. Starting in 1994, PAT was also found in experiments on single-electron transport through semiconductor quantum dots [12-16]. The quantum dots in Ref$.$ \cite{Leo-PAT2} were rather large and effectively had a continuous density of states. Here, we focus on PAT processes through quantum dots with well-resolved, discrete energy states. For these small dots, the resonant tunneling peak in the current develops photon sideband resonances when we apply microwaves \cite{Tjerk PAT}. The energy separation between main peak and sidebands can be used as a spectroscopic measurement of the energy levels in the dot. The relevant ac regime for quantum dots is at much lower frequency than visible light. We list the important frequency scales in Table I. The single-particle level spacing $\Delta \varepsilon $ is 0.05 - 0.5 meV for typical dots and the charging energy, $e^{2}/C$, is usually 0.2 - 2 meV. To observe effects from a finite $\Delta \varepsilon $ and $e^{2}/C$, these energies should exceed the thermal broadening $\sim 4k_{B}T$. Other characteristic frequencies of the dot are related to the transport times. $\Gamma $ is the typical rate to tunnel on or off the dot, which can be arbitrarily small for opaque tunnel barriers. This frequency is set by the transmission coefficient of the barriers and should be kept smaller than $\Delta \varepsilon$ otherwise the level broadening exceeds the spacing between the single-particle states. The final time scale is the tunneling time; i$.$e$.$ the actual time spent during tunneling through the barrier. This time is quite short ($\sim $2 ps) for typical barriers (calculated within the B\"{u}ttiker-Landauer framework \cite{tunnel-time}). To access these time scales, ac signals can be applied, and the effects on the dc transport can be measured. \begin{table}[htbp] \begin{tabular}[t]{|c|c|c|} \hline {\bf Quantity} & {\bf Equivalent frequency} & {\bf Typical frequencies} \\ \hline Thermal broadening & $\sim 4k_{B}T/h$ & 10 GHz (at 100 mK) \\ \hline $ \begin{array}{c} \text{Tunneling rate} \\ \text{on/off the dot} \end{array} $ & $\Gamma $ & 0 - 100 GHz \\ \hline $ \begin{array}{c} \text{Level spacing (or} \\ \text{inverse traversal time)} \end{array} $ & $\Delta \varepsilon /h$ & 10 - 100 GHz \\ \hline Charging energy & $e^{2}/hC$ & 40 - 400 GHz \\ \hline Tunneling time & $1/\tau _{tunnel}$ & 200 GHz - 1 THz \\ \hline \end{tabular} \caption{A list of the important energy/frequency scales for transport through quantum dots. For $f$ = 10 GHz the photon energy, $hf$, is 40 $\mu $eV.} \label{tab:1} \end{table} If $f$ $\ll \Gamma $ each electron sees an essentially static potential and we are in the adiabatic regime \cite{Ashoori,MarcusPump}. If $f$ $\gg \Gamma $, each electron experiences many cycles of the ac signal while it is on the dot; i$.$e$.$ the non-adiabatic regime. If $hf\ll 4k_{B}T$, single photon processes are masked by thermal fluctuations, and a classical description is appropriate \cite{Leopump}. Thus, the discreteness of the photon energy can be observed in the non-adiabatic, high-frequency regime: $hf\gg h\Gamma ,4k_{B}T$. This is the quantum, or time-dependent regime for which it is essential to solve the time-dependent Schr\"{o}dinger equation for the tunneling electron. This review is divided into two parts. First, in section \ref{theory}, we discuss the theory of PAT through a single junction (\ref{PATsing}) and calculations of PAT through a single quantum dot using a master equation approach (\ref{master}). Numerical PAT calculations are presented in sections \ref{numresults} and \ref{PAP}. Second, in section \ref{experiments}, we discuss experimental results of PAT measurements on single lateral quantum dots. The sample geometry is discussed in section \ref{sample}, the sensitivity of pumping to the applied microwave frequency in \ref{frequency}. PAT in the low ($hf<\Delta \varepsilon $) and high frequency ($hf>\Delta \varepsilon $) regime are discussed in \ref{PATlowfreq} and \ref{PAThighfreq}, respectively. In sections \ref{freqdep} through \ref{Pdep} results are given of the dependence of PAT on microwave frequency, magnetic field and microwave power, respectively. We conclude and discuss the results in section \ref{conclusions}. \section{Theory} \label{theory} \subsection{PAT through a single junction} \label{PATsing} First, we briefly outline photon assisted transport through a single tunnel junction separating two metallic leads. An oscillating potential difference across a junction, $\widetilde{V}\cos (2\pi ft)$, where $\widetilde{V}$ is the ac amplitude, may be included in the Hamiltonian of one of the leads as: $H=H_{0}+H_{ac}=H_{0}+e\widetilde{V}\cos (2\pi ft)$, where the unperturbed Hamiltonian, $H_{0}$, describes the leads without microwaves. The effect of the oscillating potential is that the time-dependent part of the electron wave function in this lead, when expanded into a power series, contains energy components at $E$, $E\pm hf$, $E\pm 2hf$, ..., etc. These are called sidebands. The expansion can be done as follows \cite{Tien}: \begin{eqnarray} \psi (r,t) &=&\varphi (r)\exp \left( -i\int dt[E+e\widetilde{V}\cos (2\pi ft)]/\hbar \right) \nonumber \\ &=&\varphi (r)\exp \left( -iEt/\hbar \right) \sum_{n=-\infty }^{\infty }J_{n}(e\widetilde{V}/hf)\exp (-in2\pi ft) \nonumber \\ &=&\varphi (r)\left( \sum_{n=-\infty }^{\infty }J_{n}(e\widetilde{V}/hf)\exp (-i[E+nhf]t/\hbar )\right) \label{wavefunction} \end{eqnarray} $\varphi (r)$ is the eigenfunction satisfying $H_{0}\varphi (r)=E\varphi (r)$ {\bf \ }and forms the spatial part of the wave function $\psi (r,t)$. $ J_{n}(\alpha )$ is the $n$th order Bessel function of the first kind (see Fig$.$ 1a) evaluated at $\alpha =e\widetilde{V}/hf$. Let $\rho _{l}$ and $\rho _{r}$ be the unperturbed densities of states of the left and right leads. The tunnel current through a junction without a microwave field is then given by \begin{equation} I(V_{SD})=c\int_{-\infty }^{\infty }dE[f_{l}(E-eV_{SD})-f_{r}(E)]\rho _{l}(E-eV_{SD})\rho _{r}(E) \end{equation} where $V_{SD}$ is the source-drain voltage, $f(E)$ is the Fermi function, and $c$ is a constant proportional to the tunnel conductance. \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig1.eps, width=12.5cm, clip=true}} \caption{(a) Squared Bessel functions of the first kind $J_n^2(\alpha)$, for $n$ = 0, 1 and 2. The inset shows the development of sidebands of the original energy as a consequence of the microwave field. The population probability $P(n)$ of the different sidebands is given by $P(n)=J_n^2(e \widetilde{V}/hf)$. A positive or negative $n$ corresponds to the absorption or emission, respectively, of $n$ photons during the tunnel process. Elastic tunneling corresponds to $n=0$. (b) Schematic energy diagram of a single dot containing $N$ electrons distributed over the available single-particle levels $\varepsilon _j$ in a particular configuration $\chi$. By absorbtion of a photon, an electron can tunnel into $\varepsilon_1$ so that the electron number changes into $N+1$ and the configuration into $\chi^{\prime}$.} \label{fig1} \end{center} \end{figure} From Eq$.$ (\ref{wavefunction}), we can write an effective density of states in one of the leads (we choose the right lead) given by: \begin{equation} \widetilde{\rho }_{r}(E)=\sum_{n=-\infty }^{\infty }\rho _{r}(E+nhf)J_{n}^{2}(e\widetilde{V}/hf) \end{equation} If tunneling is a weak perturbation, the dc current in the presence of microwaves, $\widetilde{I\text{,}}$ is given by \cite{Tien}: \begin{align} \widetilde{I}(V_{SD}) = & \hspace{0.1cm} c\sum_{n=-\infty}^{\infty}J_{n}^{2}(e\widetilde{V}/hf) \times \nonumber \\ & \int_{-\infty }^{\infty }dE[f_{l}(E-eV_{SD})-f_{r}(E+nhf)]\rho _{l}(E-eV_{SD})\rho _{r}(E+nhf) \nonumber \\ = & \sum_{n=-\infty }^{\infty }J_{n}^{2}(e\widetilde{V}/hf) I(V_{SD}+nhf/e) \end{align} We stress that tunneling is assumed to be a weak perturbation, implying that the sidebands are only well-defined for $f\gg \Gamma .$ Since there is no electric field in the scattering-free leads, mixing of electron states \cite {Platero}, or photon absorption is absent in the leads. A positive or negative $n$ corresponds to the absorption or emission, respectively, of $n$ photons during the tunnel process. Elastic tunneling corresponds to $n=0$. {\it For a single junction the dc current in the presence of microwaves is thus described, simply in terms of the dc current without microwaves.} From the normalization $\sum_{n=-\infty }^{\infty }J_{n}^{2}(\alpha )=1$, it follows that the integrated current does not change due to the oscillating potential: $\int \widetilde{I}(V_{SD})dV_{SD}=\int I(V_{SD})dV_{SD}.$ We emphasize that although the oscillating field is entirely classical, the interaction with an electron, described by the Schr\"{o}dinger equation, is only via exchange of {\it discrete }energy quanta. Equation (4) is only valid for single junctions where the tunneling takes place via a single hop. An extension of Eq$.$ (4) to describe a double junction system is discussed next. \subsection{Master equation for PAT through a quantum dot} \label{master} Electron transport through double barrier structures is resonant when the Fermi energy of the leads aligns with a discrete energy state between the two barriers. Transport through semiconductor quantum wells are usually well described by non-interacting electron models. Also, their transport properties in the presence of an oscillating signal can, in a first-order approximation, be described by the time-dependent, non-interacting Schr\"{o}dinger equation. The result of such calculations is that, next to the main resonance, extra peaks appear at distances corresponding to the photon energy, $hf$ \cite{Sokolovski,Johansson}. In quantum dots electrons are confined in all directions. The total number of electrons, and the total charge, is thus a discrete value. This makes it essential to include the Coulomb interactions when describing transport. The standard model is known as the single-electron tunneling, or the Coulomb blockade model \cite{LesHouches}. This model takes into account that at low voltages and low temperatures only one electron can tunnel at a time. The necessary energy to add an extra electron to a quantum dot consists of the charging energy $E_{c}=e^{2}/C$ for a single electron, and a discrete energy difference, $\Delta \varepsilon $, arising from the quantum-mechanical confinement. In practice, a quantum dot has discrete energy states if $ \Delta \varepsilon $ exceeds the thermal energy $k_{B}T$ \cite{Coulomb}{\bf . }Assuming sequential tunneling of single electrons, the current can be calculated with a master equation \cite{Averin,Beenakker}. PAT through small systems in which Coulomb blockade is important was considered first by Likharev and Devyatov \cite{Likharev}, Hadicke and Krech \cite{Hadicke}, and Bruder and Schoeller \cite{bruder}. A direct inclusion of the Tien and Gordon equations \cite{Tien} in a master equation that takes into account Coulomb blockade \cite{LesHouches} can be made by writing the tunnel rate through each barrier in the presence of microwaves $\widetilde{ \Gamma }(E)$ in terms of the rates without microwaves $\Gamma (E)$ \cite {Leo-PAT1}: \begin{eqnarray} \widetilde{\Gamma }(E)=\sum\limits_{n=-\infty}^{+\infty}J_{n}^{2}(\alpha )\Gamma (E+nhf) \label{PAT-rate} \end{eqnarray} Equation \ref{PAT-rate} has a direct link to studies of the effects of fluctuations in the electromagnetic environment on single-electron tunneling \cite{Ingold-Nazarov}. If the spectral density of the environment is characterized by the probability function $P(hf)$, then the rate including the environment $\Gamma _{env}(E)$, can be written in terms of the rate without the environment $\Gamma (E)$ as \cite{Hu-Connel}: \begin{eqnarray} \Gamma _{env}(E)= \int_{- \infty}^{\infty} d(hf)P(hf)\Gamma (E+hf) \label{Env-rate} \end{eqnarray} Whereas Eq$.$ (\ref{PAT-rate}) describes a monochromatic environment, the fluctuations in general are broad band in frequency, as described in Eq$.$ (\ref{Env-rate}). Examples of environments that have been studied experimentally, are the impedance in the leads \cite{Ingold-Nazarov}, blackbody radiation \cite{Harvard}, and phonons \cite{Toshi-Spont}. We note, however, that Eqs$.$ (\ref{PAT-rate},\ref{Env-rate}) are valid only for systems with a continuous density of states (i$.$e$.$ $\Delta \varepsilon $ $\ll k_{B}T$) and immediate relaxation to the ground state after each tunnel event. In the case of quantum dots with large level separation (i$.$e$.$ $\Delta \varepsilon $ $\gg k_{B}T$), one needs to keep track of the occupation probabilities of each discrete state. This increases the amount of bookkeeping, but has the advantage that intra-dot relaxation and excitation processes can be included. In our model \cite{Tjerk-PAT2} for PAT through small dots, we assume $E_{c}\gg \Delta \varepsilon $, $k_{B}T$, $eV_{SD}$, $ nhf$, such that we only need to consider two charge states (i$.$e$.$ the electron number is either $N$ or $N+1$) \cite{Wan}. We neglect level broadening due to a finite lifetime of the electrons on the dot. A charge state (Fig$.$ 1b) is described by the electron number $N$, together with the particular occupation of the electrons in the available single-particle levels $\{\varepsilon _{j}\}$. If $N$\ electrons are distributed over $k$ levels, the number of distinct dot configurations, $\chi ,$ is given by $\binom{k}{N}$. The probability, $P_{N,\chi }$, for state $(N,\chi ) $ is calculated from a set of master equations given by: \begin{eqnarray} \dot{P}_{N,\chi } &=&\sum_{\chi ^{^{\prime }}}P_{N+1,\chi ^{\prime }}(\Gamma _{l,j_{\chi ^{\prime }}}^{out}+\Gamma _{r,j_{\chi ^{\prime }}}^{out}) \nonumber \\ &&-P_{N,\chi }\sum_{j=empty}(\Gamma _{l,j}^{in}+\Gamma _{r,j}^{in}) \label{rate equation} \\ &&+\sum_{\chi ^{\prime \prime }\neq \chi }P_{N,\chi ^{\prime \prime }}\Gamma _{\chi ^{\prime \prime }\rightarrow \chi }-P_{N,\chi }\sum_{\chi ^{\prime \prime \prime }\neq \chi }\Gamma _{\chi \rightarrow \chi ^{\prime \prime \prime }} \nonumber \end{eqnarray} and the equivalent forms for $\dot{P}_{N+1,\chi ^{\prime }}$. To find a stationary solution, these equations are all set to zero ($\dot{P}=0$) and solved with the boundary condition: \begin{equation} \sum_{\chi }P_{N,\chi }+\sum_{\chi ^{\prime }}P_{N+1,\chi ^{\prime }}=1 \label{boundary condition} \end{equation} For $N=2$ distributed over five different single-particle levels $\{\varepsilon _{j}\}$, there are ten different configurations, $\chi $, yielding ten equations for $\dot{P}_{N,\chi }$ and also ten equations for $\dot{P}_{N+1,\chi ^{\prime }}$. The first and second term in Eq$.$ (\ref{rate equation}) correspond to a change in the occupation probability of a certain distribution due to tunneling (the number of electrons on the dot changes). In the first term an electron tunnels out of the dot. Only those rates are taken into account that correspond to an electron tunneling out of state $j_{\chi ^{\prime }}$ that leave the dot in the distribution ($N$, $\chi $). In the second term, an electron tunnels onto the dot. One needs to sum over all the states $j$ that are empty when the dot is in configuration $\chi $, because all these events cause a transition from state $(N,\chi )$ to a state $(N+1,\chi ^{\prime })$. $\Gamma _{l,j}^{in}\ $and $\Gamma _{l,j}^{out}$ are the tunnel rates through the left barrier in and out of single-particle level $j$ on the dot: \begin{eqnarray} \Gamma _{l/r,j}^{in}(\varepsilon _{j}) &=&\Gamma _{l/r,j}\sum_{n}J_{n}^{2}(\alpha _{l/r})f(\varepsilon _{j}-\frac{C_g}{C}eV_{g}-nhf+\eta _{l/r}eV_{SD};T_{l/r}) \label{rates} \\ \Gamma _{l/r,j}^{out}(\varepsilon _{j}) &=&\Gamma _{l/r,j}\sum_{n}J_{n}^{2}(\alpha _{l/r})[1-f(\varepsilon _{j}-\frac{C_g}{C}eV_{g}-nhf+\eta _{l/r}eV_{SD};T_{l/r})] \nonumber \end{eqnarray} where $\Gamma _{l/r,j}$ is the tunnel rate through the left or right barrier of energy level $j$. $\alpha _{l/r}$ is the parameter describing the microwave field at the left or right barrier, $C_{g}$ is the gate capacitance, $C$ is the total dot capacitance, $T_{l/r}$ is the temperature of the left or right lead. $\eta _{l/r}$ is a parameter describing the asymmetry of the dc voltage drop across the two barriers. We assume that the tunnel rates, $\Gamma _{l/r,j}$, do depend on the level index $j$, but that they are independent of energy. In the last two terms of Eq$.$ (\ref{rate equation}), the number of electrons on the dot is fixed, while only the distribution of the electrons over the states changes. This includes effects from relaxation (i$.$e$.$ intra-dot transitions, $\chi ^{\prime \prime }\rightarrow \chi $ and $\chi \rightarrow \chi ^{\prime \prime \prime },$ where the total energy decreases) or excitation inside the dot (i$.$e$.$ intra-dot transitions, $\chi ^{\prime \prime }\rightarrow \chi $ and $\chi \rightarrow \chi ^{\prime \prime \prime },$ where the total energy increases). Below, we take excitation rates equal to zero (i$.$e$.$ no intra-dot absorption) but, allow for non-zero relaxation rates. An expression for the dc current can be found by calculating the net tunnel rate through one of the barriers. Using the probabilities $P_{N,\chi }$ and the tunnel rates through the left barrier, this leads to: \begin{equation} I=e\sum_{\chi }\sum_{j=empty}P_{N,\chi }\Gamma _{l,j}^{in}-e\sum_{\chi ^{\prime }}\sum_{j=full}P_{N+1,\chi ^{\prime }}\Gamma _{l,j}^{out} \end{equation} {\bf \ } In the numerical calculations in the next section we take equal ac amplitudes dropping across the left and right barriers; i$.$e$.$ $\alpha _{l}=\alpha _{r}=\alpha $. \subsection{Numerical results} \label{numresults} Figure \ref{fig2} shows calculations without relaxation between the states in the dot. The inset shows the case for transport through only a single level; $\Delta \varepsilon \gg hf$. Next to the main resonance, side-peaks develop at multiples of $hf/e$ when the microwave power is increased via the parameter $\alpha .$ The broadening of the resonances is due to a finite temperature. In the main figure transport can also occur via excited states; here $\Delta \varepsilon < hf$. Not only side-peaks develop, but also peaks at other gate voltages. These peaks arise due to the interplay between the discrete single-particle states and the photon energy. Their locations are given by $(m\Delta \varepsilon +nhf)/e$ where $m=0,$ $ \pm 1,$ $\pm 2,$ ... and $n$ is the photon number. Similar simulation results have been reported by Bruder and Schoeller \cite{bruder}. \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig2.eps, width=10.5cm, clip=true}} \caption{Calculation without relaxation. Curves are offset for clarity. The parameters for the data in the inset are $\Delta \varepsilon =3hf$, $hf=5k_{B}T$, and from top to bottom $\alpha =0$, 1, 1.5, 2. The parameters for the main figure are $ \Delta \varepsilon =0.75hf$, $hf=20k_BT$\ and from top to bottom $\alpha =0 $, 0.5, 0.75, 1, 1.5.} \label{fig2} \end{center} \end{figure} Figure \ref{fig3} shows an expansion for the curve with $\alpha =1$. We have assigned the excited states and the particular PAT processes. The highest occupied single-particle level of the $N+1$ ground state is denoted by $\varepsilon _{j}$ with $j=0$; positive $j$'s are excited states above $ \varepsilon _{0}$ and negative $j$'s are below $\varepsilon _{0}$ (see Fig$.$ 1b). The inset shows the effect of relaxation. Upon increasing the relaxation rate, the peaks corresponding to tunneling through excited states decrease, while peaks increase when tunneling occurs through the ground state. \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig3.eps, width=10.5cm, clip=true}} \caption{Expansion for the curve $\alpha =1$ from Fig$.$ 2. The inset shows the effect of an increasing relaxation rate. The relaxation rates divided by the tunnel rate are 0, 0.1, 0.35, 1, 3.5, and infinite.} \label{fig3} \end{center} \end{figure} \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig4.eps, width=10cm, clip=true}} \caption{Diagrams depicting the tunneling events which dominantly contribute to the current through a quantum dot at different gate voltages. A small dc bias raises the left Fermi level with respect to the right Fermi level. $\varepsilon _{0}$ and $\varepsilon _{1}$ denote the ground state and the first excited state of the $(N+1)$ electron system. When the $(N+1)^{th}$ electron tunnels to one of the two reservoirs, the energy states of the dot drop by the charging energy $ E_{c}$. The corresponding diagrams for $N$ electrons are not shown.} \label{fig4} \end{center} \end{figure} To explain these numerical results, we show energy diagrams in Fig$.$ \ref{fig4}, assuming that only the highest two single-particle levels contribute to the current for the transition between $N$ and $N+1$ electrons on the dot. For small dc bias voltage and no ac voltages a current resonance occurs when the topmost energy state (i$.$e$.$ the electrochemical potential) of the quantum dot lines up with the Fermi levels of the leads (see the diagram $\varepsilon _{0}$). When high-frequency voltages drop across the two barriers, additional current peaks appear. We distinguish two mechanisms. The first mechanism gives photon induced current peaks when the {\it separation} between the ground state $\varepsilon _{0}$ and the Fermi levels of the leads {\it matches} the photon energy (or multiples, $nhf$), as depicted in the diagrams labeled by $\varepsilon _{0}+hf$ and $\varepsilon _{0}-hf$. The minus and plus signs correspond to being before or beyond the main resonance. Note that also the case of $\varepsilon _{0}-hf$ involves photon absorption. Following the literature on the tunneling time, we call these current peaks: {\it sidebands} \cite{tunnel-time}. The second mechanism leads to photon peaks when an excited state is in resonance with the Fermi levels of the leads (see diagram $\varepsilon _{1}$). Without PAT, transport through the excited state, $\varepsilon _{1}$, is blocked since Coulomb blockade prevents having electrons in both the ground state and the excited state simultaneously. The electron in the ground state cannot escape from the dot, because its energy is lower than the Fermi levels in the leads. PAT, however, can empty the ground state $\varepsilon _{0}$ when the electron absorbs enough energy and leaves the dot. This process is analogous to photo-ionization. Now, the ($N+1)^{\text{th}}$ electron can tunnel resonantly via the excited state $\varepsilon _{1}$ as long as the state $ \varepsilon _{0}$ stays empty. Note that for this second mechanism $nhf$ has to {\it exceed}, but not necessarily {\it match} the energy splitting $ \Delta \varepsilon =\varepsilon _{1}-\varepsilon _{0}$. It is clear from these diagrams that relaxation from $\varepsilon _{1}$ to $\varepsilon _{0}$ decreases the height of this resonant peak. More photon peaks are generated when these two mechanisms are combined as in the diagrams labeled by $ \varepsilon _{1}+hf$ and $\varepsilon _{1}-hf$. We thus see that PAT can populate the excited states by tunneling between dot and leads. So, even without intra-dot transitions, we can perform photon spectroscopy on discrete quantum dot states. \subsection{Photon-assisted pumping} \label{PAP} It is important to note that in the diagrams of Fig$.$ \ref{fig4} only processes with tunneling from or to states in the leads close to the Fermi levels contribute to the net current. Tunnel processes that start with an electron in one of the leads from further below the Fermi level are cancelled by an electron from \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig5.eps, width=9cm, clip=true}} \caption{Calculation of the current for zero-bias voltage $(V=0)$\ as a function of gate voltage in the case where the ac voltage drop over one barrier is 5\% smaller than over the other barrier $(k_{B}T=0.05hf$, $\Gamma =5 10^{8}$\ s$^{-1})$. The insets depict which tunneling events are responsible for the pumped current when the ground state of the quantum dot is below or above the Fermi levels of the leads.} \label{fig5} \end{center} \end{figure} the other lead. However, this is only true when the ac voltage drop is the same for both barriers. When the ac voltage drops across the two barriers are unequal, the dot acts as an electron pump \cite{Leopump,Leo-PAT1,bruder}. The resulting pumped current makes the resonances discussed above less clear. For this reason we discuss this pumping mechanism in more detail here before proceeding further. Figure \ref{fig5} shows a calculation of the pumped current as a function of the gate voltage that occurs when the ac voltage drop over one barrier is 5\% smaller than over the other barrier. We have taken zero dc bias voltage. To illustrate the origin of the pumped current, the insets show the extreme case, when all the ac voltage drop is across the left barrier. In this case photon absorption occurs only at the left barrier. At negative gate voltage, when the ground-state level of the dot is {\it above} the Fermi level of the leads, an electron can {\it enter} the dot from the left lead only (bottom left inset to Fig$.$ \ref{fig5}). Once the electron is in the dot, it can tunnel out through both tunnel barriers. Only tunneling to the right lead contributes to the net current. Therefore, the net (particle) current is to the right. When the ground-state level of the dot is {\it below} the Fermi level of the leads, however, an electron can only {\it leave} the dot to the left lead (upper right inset to Fig$.$ \ref{fig5}). The dot can be filled from either lead once it is emptied. This time only the electron tunneling {\it in} from the right lead contributes to the net current. Therefore, there is a net current to the left. The difference between these two situations is the shift in the ground-state energy with respect to the Fermi levels of the leads. So, when the gate voltage is swept such that the ground state moves through these Fermi levels, the pumped current changes sign. The pumped current occurs over a width corresponding to the photon energy. The extra shoulders at the far left and far right are due to two-photon processes. \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig6.eps, width=9cm, clip=true}} \caption{Simulation of the effects of both asymmetric heating and quantum rectification on the curve with $\alpha =1$ from Fig$.$ 3. The solid line represents the unperturbed curve. The dashed-dotted line shows the same curve for $T_{R}/T_{L}$ = 0.95. The dashed line shows the influence of $\alpha _{R}/\alpha _{L}$ = 0.95 on the unperturbed curve.} \label{fig6} \end{center} \end{figure} Asymmetric heating may induce a difference in the temperatures $T_{R}$ and $ T_{L}$ of the two leads. This can also result in a finite transport current. The effects of asymmetric heating and asymmetric coupling of the microwave signal when a finite $V_{SD}$ is applied across the sample, is illustrated in Fig$.$ \ref{fig6}. The solid line is a reproduction of Fig$.$ \ref{fig3}. The dashed-dotted line shows the same trace for $T_{R}/T_{L}$ = 0.95. The dashed line shows the influence of $\alpha _{R}/\alpha _{L}=0.95$. It can be concluded that both effects can severely distort the data. At small $V_{SD}$ the resonant peaks scale linearly with $V_{SD}$, while the pumped current does not change. Increasing the bias, while remaining in the linear regime, therefore improves the visibility of the resonant peaks. \section{Experiments} \label{experiments} \subsection{Sample geometry} \label{sample} \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig7.eps, width=9.5cm, clip=true}} \caption{SEM photo of the sample. The lithographic size of the dot is ($600\times 300$) nm$^{2}$. Current can flow when we apply a voltage between source and drain. The microwave signal is capacitively coupled to one of the center gates.} \label{fig7} \end{center} \end{figure} Our measurements are performed on a quantum dot defined by metallic gates (see Fig$.$ \ref{fig7}) in a GaAs/AlGaAs heterostructure containing a 2-dimensional electron gas (2DEG) 100 nm below the surface. The 2DEG has mobility 2.3 10$^{6}$ cm$^{2}$/Vs and electron density 1.9 10$^{15}$ m$^{-2}$ at 4.2 K. By applying negative voltages to the two outer pairs of gates, we form two quantum point contacts (QPCs). An additional pair of center gates between the QPCs confines the electron gas to a small dot. No electron transport is possible through the narrow channels between the center gates and the gates forming the QPCs. The center gate voltage, $V_{g}$, can shift the states in the dot with respect to the Fermi levels of the leads and thereby controls the number of electrons in the dot. The energy shift is given by $\Delta E=\kappa \Delta V_{g}$, with $\kappa$ defined as the ratio between the dot-gate capacitance and the total capacitance of the dot \cite{Coulomb}. A small dc voltage bias is applied between source and drain and the resulting dc source-drain current is measured. From standard dc measurements we find that the effective electron temperature is approximately $T$ = 200 mK and the charging energy $E_{c}=1$.2$ \pm $0.1 meV. We independently determine the level splitting, $\Delta \varepsilon$, for different magnetic fields from current-voltage characteristics. In addition to the dc gate voltages, we couple a microwave signal (10-75 GHz) capacitively into one of the center gates. The microwave does not equally couple to the dot as to the leads, which results in an ac voltage drop over both barriers. \subsection{Frequency sensitivity of pumping} \label{frequency} We first present experimental results with a strongly pumped current taken at $B$ = 1.96 T for three frequencies around 47.4 GHz (the arrow denotes $hf$). The dashed line in Fig$.$ \ref{fig8} is the current without microwaves. For the lowest frequency the \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig8.eps, width=12cm, clip=true}} \caption{Measurements of the pumped current at $B$ = 1.96 T, $V_{SD}$ = 13 $\mu$V and frequencies are around 47.4 GHz. Dashed line is without microwaves. The dotted line shows the smallest asymmetry, but shows evidence for a pumping mechanism which is not included in our model.} \label{fig8} \end{center} \end{figure} current is pumped in one direction, whereas for the highest frequency it is pumped in the opposite direction. At 47.33 GHz the left barrier apparently has the {\it smaller} ac voltage drop, while at 47.43 GHz the left barrier has the {\it larger} ac voltage drop. This illustrates that the asymmetry of the voltage drops over the two barriers sensitively depends on frequency. This sensitivity is ascribed to standing waves in the sample holder. The dotted line shows the current measured at an intermediate frequency, where we expect the ac voltage drop to be equal over both barriers. In contrast to the two solid curves, the dotted line is lower than the dashed line without microwaves over the whole gate voltage range. This cannot be explained by the pumping mechanism in our model. Our model only includes the oscillation of the potential of the leads relative to the dot, which always results in a pumped current which changes sign at the resonance. A negative pumped current over the whole gate voltage range, is attributed to the effect of the microwaves on the barrier height. The inset shows how a quantum dot can act as a pump when one tunnel barrier is periodically modulated in height. During one part of the cycle, when the left barrier is low, electrons enter the dot ($\Gamma _{L}^{low}>\Gamma _{R}$ ) while they escape the dot through the right barrier in the second half of the cycle when the left barrier is high ($\Gamma _{L}^{high}<\Gamma _{R}$). This is essentially a classical mechanism that has been verified experimentally in the MHz regime \cite{Leopump}. For observing clearly separated PAT sidebands, we first minimize pumping. For this we measure traces of current at zero bias voltage across a single Coulomb peak for slightly different frequencies. We finally choose those frequency values for which the pumped current is very small. \subsection{PAT: Low frequency regime} \label{PATlowfreq} First, we study the photon sidebands of the ground state at $B=0$.84 T \cite {B not zero}. The main part of Fig$.$ \ref{fig9} shows measured curves of the current as a function of the gate voltage at different microwave powers for the case $hf<\Delta \varepsilon $. Here, current flows primarily via the ground state and its photon sidebands (i$.$e$.$ upper diagrams in Fig$.$ \ref{fig4}). On increasing the microwave power, we see that the height of the main resonance decreases to zero while additional resonances develop with increasing amplitude. When we convert gate voltage to energy, we find that the additional resonances are located at $\varepsilon _{0}\pm hf$ and $ \varepsilon _{0}\pm 2hf$ \cite{2hf>Deltae}. The power dependence is in agreement with the behavior of the Bessel functions: $J_{0}^{2}(\alpha )$ for the main resonance $\varepsilon _{0}$, $J_{1}^{2}(\alpha )$ for the one-photon sidebands $\varepsilon _{0}\pm hf$, and $J_{2}^{2}(\alpha )$ for the two-photon sidebands $\varepsilon _{0}\pm 2hf$. For comparison, we show a calculation in the inset to Fig$.$ \ref{fig9} for the same values for the temperature, frequency and bias voltage as in the experiment. We have assumed equal ac voltages across the two barriers. The difference between measured and calculated data is attributed to an asymmetry in the ac coupling. \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig9.eps, width=11cm, clip=true}} \caption{Measurement of the current through the quantum dot as a function of the center gate voltage and the output voltage of the microwave supply. These data are taken in the single-level regime ($hf < \Delta \varepsilon$). $hf$ = 110 $\mu$eV for $f$ = 27 GHz, $\Delta \varepsilon$ = 165 $\mu$eV at $B$ = 0.84 T, and $V_{SD}$ = 13 $\mu$V. Inset: calculation of the current as a function of the gate voltage and the ac voltage parameter $\alpha =e\tilde{V}/hf$, taking the same values for $T$, $f$, and $V$ as in the experiment.} \label{fig9} \end{center} \end{figure} \subsection{PAT: High frequency regime} \label{PAThighfreq} \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig10.eps, width=11cm, clip=true}} \caption{Measured current as a function of center gate voltage for different microwave powers. The dashed curve is without microwaves. $B$ = 0.91 T, $V_{SD}$ = 13 $\mu$V. $f$ = 61.5 GHz in the top section, $f$ = 42 GHz in the bottom section. As the frequency is reduced between top and bottom sections, the ground-state resonance $\varepsilon _0$ and the resonance attributed to the excited state $\varepsilon _1$ remain at the same gate voltage position. The other peaks, $\varepsilon _0-hf$ and $\varepsilon_1 \pm hf$, shift inward by an amount which corresponds to the change in photon energy as indicated by the arrows. We do not observe $\varepsilon _{0}+hf$\ in this measurement.} \label{fig10} \end{center} \end{figure} \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig11.eps, width=8cm, clip=true}} \caption{$I_{DC}-V_{g}$ curves for increasing frequency; $B$ = 0.91 T, $V_{SD}$ = 13 $\mu$V and powers range between -30 and 0 dBm. The dashed curve is without microwave power. The dotted curve is with maximum power. Five different peaks can be distinguished in the data. These peaks, labelled A through E for the 61.45 GHz traces, correspond to the $\varepsilon_0 - hf$, $\varepsilon_1 - hf$, $\varepsilon_0$, $\varepsilon_1$ and $\varepsilon_{1}+hf$ peaks, respectively.} \label{fig11} \end{center} \end{figure} We now discuss the higher frequency regime where $hf>\Delta \varepsilon$, such that PAT can induce current through excited states. Figure \ref{fig10} shows the current at $B$ = 0.91 T (here $\Delta \varepsilon =130$ $\mu $eV). In the top section $f$ = 61.5 GHz ($hf$ = 250 $\mu$eV) and in the bottom section $f$ = 42 GHz ($hf$ = 170 $\mu$eV). As we increase the power, we see extra peaks coming up. We label the peaks as in Fig$.$ \ref{fig4}. On the right side of the main resonance a new peak appears, which we assign to photo-ionization, followed by tunneling through the first excited state. At higher powers the one-photon sidebands of the main resonance as well as those of the excited state resonance appear. We do not observe the peak for $\varepsilon_{0}+hf$, in this measurement. This can be explained, at least in part, by the fact that here an electron can also tunnel into $\varepsilon _{1}$, which blocks the photon current through $ \varepsilon _{0}+hf$. Simulations confirm that the peak for $\varepsilon _{0}+hf$ can be several times weaker than the peak for $\varepsilon _{0}-hf$ \cite{Tjerk-PAT2}. Also, it is masked by the high peak for $\varepsilon _{1}$ right next to it. The arrows underneath the curves mark the photon energy. The peaks $\varepsilon _{0}$ and $\varepsilon _{1}$ remain in place when we change the frequency, since the photon energy evidently does not alter the energy splitting. The other peaks, $\varepsilon _{0}-hf$ and $\varepsilon _{1}\pm hf$, shift by an amount that corresponds to the change in photon energy as indicated by the arrows. This reflects that the sidebands originate from matching the states $\varepsilon _{0}$ and $\varepsilon _{1}$ to the Fermi levels of the leads by a photon energy $hf$. Figure \ref{fig11} shows a large data set. In each panel, different traces are taken at different microwave power. The panels differ in frequency. We further substantiate the peak assignment below by studying detailed frequency, magnetic field and power dependence. \subsection{Frequency dependence} \label{freqdep} Figure \ref{fig12} shows the spacing between a resonance and its photon sidebands as a function of the photon energy. Different markers correspond to different photon sidebands. The factor $\kappa =3$5 $\mu eV/mV$, to convert the peak spacings in mV gate voltage into energy, is determined from dc measurements. The full width at half maximum (FWHM) of the resonance without microwaves, indicated by the arrow, is proportional to the effective electron temperature in the leads. Structure due to photon energies below this value is washed out by the thermal energy $k_{B}T$. The frequency scaling firmly establishes PAT as the transport mechanism [9,11-14]. The observation that the sidebands move linearly with frequency, while the ground and excited state resonances stay fixed, supports our identification of the different peaks. \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig12.eps, width=10cm, clip=true}} \caption{Peak spacings versus photon energy. $\square$: spacing between $\varepsilon_{0}$ and $\varepsilon _{0}-hf $. $\lozenge$: spacing between $\varepsilon_{0}$ and $\varepsilon _{0}+hf$. $\triangle $: spacing between $\varepsilon_1$ and $\varepsilon_1-hf$. {\large $\circ$}: spacing between $\varepsilon_1$ and $\varepsilon_1+hf$. The dashed line is based on the gate voltage to energy conversion factor $\kappa$ determined independently from dc measurements, and has the theoretically expected slope equal to 1. The arrow indicates the FWHM of the main resonance.} \label{fig12} \end{center} \end{figure} \subsection{Magnetic field dependence} \label{Bdep} \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig13.eps, width=8.5cm, clip=true}} \caption{(a) Peak positions in gate voltage versus magnetic field at 52.5 GHz. Solid symbols denote peaks which are independent of frequency. Open symbols denote peaks that scale with frequency. (b) Peak spacings relative to the main resonance converted to energy. Closed circles: $\varepsilon_{0}$; open circles: $\varepsilon_{0} \pm hf$; closed diamonds: $\varepsilon_{1}$; open diamonds: $\varepsilon_{1}-hf$\ and $\varepsilon _{1}-2hf$.} \label{fig13} \end{center} \end{figure} We now use a magnetic field to change the energy separation between the ground state and the first excited state \cite{Coulomb,Johnson}, while keeping the distance to the sidebands fixed. Figure \ref{fig13}a shows the positions in gate voltage of all observed peaks for 52.5 GHz as a function of magnetic field. The filled circles reflect the evolution of $\varepsilon _{0}$ with magnetic field. This ground state weakly oscillates with a periodicity of $\sim $80 mT which roughly corresponds to the addition of an extra flux quantum to the dot. The filled diamonds reflect the evolution of $ \varepsilon _{1}$. The open circles (diamonds) show the sidebands $ \varepsilon _{0}\pm hf$ ($\varepsilon _{1}\pm hf$). Figure \ref{fig13}b shows the magnetic field evolution of the excited state and the photon sideband peaks relative to the ground state (i$.$e$.$ we have subtracted $ \varepsilon _{0}(B)$ from the other curves). We see that the energy splitting decreases on increasing the magnetic field and for $0.54$ T $ <B<0.58$ T a degeneracy of the ground state is temporarily lifted and actually two excited states are observed \cite{fielddependence}. The dashed lines denote the photon energy $hf$ = 217 $\mu$eV for 52.5 GHz. The open circles close to these lines are the photon processes $\varepsilon_{0}\pm hf $, demonstrating that they indeed move together with the ground state. The open diamonds are the $\varepsilon_{1}-hf$ and $\varepsilon _{1}-2hf$ processes. Their motion follows the motion of $\varepsilon _{1}$. We have thus shown that we can vary the states $\varepsilon _{0}$ and $\varepsilon _{1}$ with the magnetic field and, independently, vary the separation to the sidebands with the microwave frequency. \subsection{Power dependence} \label{Pdep} \begin{figure}[htbp] \begin{center} \centerline{\epsfig{file=fig14.eps, width=8.5cm, clip=true}} \caption{(a) Calculation of the peak heights as a function of the ac voltage drop across the barriers $\alpha =\frac{e\widetilde{V}}{hf}$ ($T$ = 200 mK, $V$ = 13 $\mu$V, $f$ = 52.5 GHz). The tunnel rates from the leads to the ground state and the excited state are set to $\Gamma_{\varepsilon_0}$ = 5 10$^{8}$ s$^{-1}$ and $\Gamma_{\varepsilon_1}$ = 14 10$^{8}$ s$^{-1}$, respectively. The relaxation rate from the excited state to the ground state is assumed to be zero in the calculation. (b) Experimentally obtained peak heights as a function of the ac voltage amplitude (measured at the microwave source) for $V$ = 13 $\mu$V and $f$ = 52.5 GHz.} \label{fig14} \end{center} \end{figure} Figure \ref{fig14}a shows a calculation of the peak heights as a function of the ac voltage drop across the barriers $\alpha =\frac{e \widetilde{V}}{hf}$. Temperature, source-drain voltage, and frequency are taken from the experiment described below: $T=200$ mK, $V_{SD}=13$ $\mu $V, and $f=52.5$ GHz. The tunnel rates from the leads to the ground state and the excited state are set to $\Gamma_{\varepsilon_0}$ = 5 10$^{8}$ s$^{-1}$ and $\Gamma_{\varepsilon_1}$ = 14 10$^{8}$ s$^{-1}$, respectively. The relaxation rate from the excited state to the ground state is assumed to be zero in the calculation. The effect of a finite relaxation rate is to reduce the height of $\varepsilon_{1}$ with respect to the other peaks. The calculated peak heights roughly follow the Bessel functions in Eq$.$ (\ref{rates}). The ground-state resonance $\varepsilon_{0}$ follows $J_{0}^{2}(\alpha )$, since it involves only elastic tunnel events (see Fig$.$ \ref{fig4}, diagram $\varepsilon _{0}$). The photon sidebands follow $J_{1}^{2}(\alpha )$, since they solely depend on the probability of photon absorption. For example, the process $\varepsilon _{0}-hf$ is due to a photon assisted tunnel event which fills the dot. Once the dot is filled, however, it does not matter whether the dot is emptied via an elastic or an inelastic event. The process $\varepsilon _{1}$ follows the product of the Bessel functions $J_{0}^{2}(\alpha )J_{1}^{2}(\alpha )$ since it requires that the ground state is emptied via a PAT process, but also that the following tunneling processes through the excited state $ \varepsilon _{1}$ are elastic. Figure \ref{fig14}b shows the experimental results for the peak heights at $B$ = 0.91 T and $f$ = 52.5 GHz as a function of the ac voltage amplitude at the output of the source. The measurements are in good qualitative agreement below an ac source voltage of 100 mV. At higher ac voltages the pumped current starts to become important. The values for the tunnel rates to $ \varepsilon _{0}$ and to $\varepsilon _{1}$ derived from the dc current-voltage characteristic are $\Gamma _{\varepsilon_0}$ = 5 10$^{8}$ s$^{-1}$ and $\Gamma_{\varepsilon_1}$ = 6 10$^{8}$ s$^{-1}$. The value for $\Gamma_{\varepsilon _{1}}$ in the calculation is larger than the experimentally determined value, but still the calculated value for the height of the $\varepsilon _1$ resonance is smaller than the experimental value. It is a general trend in most of our data that the peak $ \varepsilon _{1}$ is higher than predicted by our model and that $ \varepsilon _{0}+hf$ is lower than expected from simulations. A possible explanation has been put forward in Ref$.$ \cite{Brune}. \subsection{Discussion and Conclusions} \label{conclusions} The simulations described earlier, show that the pumped current is quite independent of the bias voltage when $eV_{SD}\ll hf$, while current due to the photon resonances increases linearly with the bias voltage when $ eV_{SD}<k_{B}T$. Therefore, it is possible to improve the quality of our data by separating the pumped current from the photon resonances. This could be done by repeating a measurement at a particular microwave power for different bias voltages. In addition, the effective electron temperature of 200 mK reported here has now been improved to 50 mK. These two improvements would allow for a better comparison of future experiments with calculations over a wider range of microwave powers. The conclusion of this work is that photon assisted tunneling is clearly observed in single-electron transport through small quantum dots. In addition, microwave irradiation can be used to perform spectroscopy on the discrete level spectrum. The parameter dependence of PAT is in reasonable agreement with calculations based on a master equation. Both, linear-frequency dependence and Bessel function power dependence are clearly observed. Recently, a non-equilibrium Green's function method applied to our PAT studies, has provided very good agreement with our results, including an explanation for the absence of the sideband at $\varepsilon _{0}+hf$ \cite{Sun}. Brune {\it et al}. \cite{Brune} have analyzed the influence of intra-dot transitions. Including intra-dot transitions, they find good agreement with the height of the measured $\varepsilon_{1}$ peak in Figs$.$ \ref{fig10} and \ref{fig11}. Qualitatively similar results have been obtained on superlattice structures irradiated by a free-electron laser. The dot and superlattice experiments have stimulated new theoretical interest in ac transport through non-superconducting structures. For instance, new results have been obtained for PAT across a single tunnel barrier for different sorts of oscillating potentials \cite{Wagner}; for PAT through superlattices in a self-consistent treatment \cite{Aguado}, for the ac effects on transport through a QPC and double barrier structures in a numerical study \cite{Yakubo}, for the ac effects on correlated transport through Luttinger liquids \cite{Cuniberti}, and for the Kondo effect in irradiated quantum dots \cite{Lopez}. These models and theories are very useful in explaining and predicting new transport mechanisms. We feel, however, that a word of caution is appropriate here. Although our PAT experiments give clean results, we have very little control over the oscillating potentials inside the sample. We simply couple in a microwave signal to one of the gates and measure the effect in the dc current. The dc response is very sensitive to the applied frequency. Sometimes the microwaves couple in more across the left barrier and sometimes more across the right barrier. The sensitivity of the asymmetry in coupling might be due to a complicated electric field pattern around the metallic gate structure. We think that to some degree our gate structure acts as a co-planar waveguide where the two QPC barrier gates serve as ground planes for the central microwave gate. Even under such conditions, it is not clear how this oscillating gate potential is carried over to the electron gas 100 nm below the surface. Pedersen and B\"{u}ttiker \cite{Pedersen} have recently analyzed this problem for the case of quantum dots. They stress that oscillating potentials not only wiggle energy levels but also generate alternating currents in the sample. The alternating currents affect the self-consistent potential of the tunnel barriers and the potential on the dot. In effect, all the voltages and capacitances become renormalized and depend on the occupation of the dot. Sideband positions and heights are not expected to precisely follow a linear frequency dependence nor a Bessel function behavior. An experimental, quantitative study of the parameter dependence of the sidebands might give valuable information about these screening effects. Even more interesting would be to study the frequency dependence of the current, including measurements of higher harmonics at multiples of $f$. An analysis for such a set-up in terms of a frequency dependent environment is given in Ref$.$ \cite{Aguado2}. PAT is intrinsically a coherent phenomenon. The PAT measurements described above, however, are insensitive to the phase of the transmitted electrons. Coherence in the presence of a time-dependent field is therefore not directly demonstrated. Jauho and Wingreen \cite{Jauho} have proposed a PAT measurement through a quantum dot situated in one of two branches of an Aharonov-Bohm ring. They find that coherent absorption and reemission of photons can be detected via a phase measurement at the sidebands. The proposed mesoscopic double-slit geometry has been successfully used before to demonstrate coherent transmission through a quantum dot for the time-{\it independent} case \cite{Yacoby,Schuster}. By connecting two quantum dots in series, a double quantum dot system can be obtained. An overview of charging effects in double quantum dots is given in Ref$.$ \cite{Coulomb}. Theoretical studies on PAT in a double quantum dot are done by Stoof and Nazarov \cite{Stoof}, Stafford and Wingreen \cite{Stafford} and Brune {\it et al}. \cite{Brune1}. PAT studies of a double quantum dot enable the characterization of the coupling between two discrete energy levels. Depending on the strength of the inter-dot coupling, the two dots can form `ionic' or `covalent' bonds. By varying the inter-dot coupling, Oosterkamp {\it et al. } \cite{Oosterkamp} experimentally demonstrated the transition from ionic bonding to covalent bonding in a quantum dot 'artificial molecule' that is probed by microwave excitations. In the same frequency regime experiments have been performed by Blick {\it et al.} \cite {Blick96} and by Fujisawa and Tarucha \cite{Toshi1}. The study of photon assisted tunneling in quantum dots described here, forms a valuable extension of the understanding of the dc properties. Microwave measurements have been proved to be a useful spectroscopy tool for quantum dot systems. Microwaves are also expected to play a role of importance in future experiments on quantum dots and the possible application of quantum dots as solid state quantum bits. An example would be a time-resolved measurement of Rabi oscillations in a double quantum dot. In this case, the microwave signal is used to tune the Rabi oscillation frequency \cite {Stoof,Stafford}. We thank R. Aguado, S.F. Godijn, A.E.A. Koolen, J.E.\ Mooij, Yu.V. Nazarov, R.M. Schouten, T. Fujisawa, S. Tarucha, T.H. Stoof, P. McEuen and N.C. van der Vaart for experimental help and useful discussions. This work was supported by the Dutch Organization for Research on Matter (FOM), by the EU via the TMR network (ERBFMRX CT98-0180), and by the NEDO joint research program (NTDP-98).
\section{Introduction} The evolution of disc galaxies is driven by both internal and external processes. Internal instabilities in the disc often give rise to the formation of a bar. Bars are a common feature in disc galaxies and by no means an exception, since about $2/3$ of all disc galaxies are believed to harbor a bar or oval distortion. Furthermore in the last few years near-infrared observations have given clear indications of an even higher fraction of barred galaxies. The presence of a bar changes both the kinematics and mass (stellar and gas) distributions within the disc and can give rise to dynamical resonances (e.g., \opencite{Sel93}). Similarly, it has become increasingly recognized that interactions with small companions are a frequent and important external agent driving dynamical evolution. Since both these processes are thought to be frequent, it is a reasonable assumption that the interaction between a barred disc galaxy and a small companion is a common event. The study of \inlinecite{Ath97}, using pure N-body simulations, has demonstrated that a sufficiently massive companion hitting the inner parts of a barred disc galaxy can displace the bar and produce expanding stellar rings. In this work we have included the dissipative effects of a gas component. \section{Numerical Methods} For the simulations we use an N-body/SPH code to evolve the stellar and gaseous components of the galaxies. A detailed description of the algorithm can be found in \inlinecite{Hel95} and \inlinecite{HeS94}. The gravitational forces and neighbor interaction lists were computed using the special purpose hardware GRAPE-3AF. The advantage of the GRAPE hardware, besides its speed, is that it does not set any constraints on the symmetry of the simulations. The adopted units for mass, distance, and time are $M\!=\!6\cdot10^{10}$\,M$_{\odot}$, $R\!=\!3$\,kpc, and $\tau\!=\!10^7$\,yr, respectively, and the gravitational constant is taken as unity. A fixed gravitational softening length of $\epsilon\!=\!0.1875$\,kpc is used for all particles. \subsection{Initial Conditions} Both the stellar and gaseous discs are initially setup with a truncated Kuzmin-Toomre projected radial density profile and a $sech^2$ vertical distribution. The halo and companion have a truncated Plummer profile. The parameters are given in Table\,1 and were chosen so as to give a bar unstable model. Based on the potential of the disc and relaxed halo we assign velocities to the disc particles. The velocity dispersion is derived from the local stability criterion and is corrected for asymmetric drift. The resulting rotation curve has a maximum of $v_{\rm max}\!\approx\!200$\,km/s and remains flat in the outer disc. The initial conditions for the interactions are specified by the impact position ($x_{\rm i},y_{\rm i}$) with respect to the bar and the impact time $t_{\rm i}$. The impact velocity $v_{\rm i}$ of the companion is chosen to be roughly equal to $4\,v_{\rm esc}$, with $v_{\rm esc}$ being the escape velocity from the center of the target system. The orbit is integrated backward for a time period of $\Delta t\!=\!30$, which places the companion well outside the halo of the disc system. A more complete description of the procedure for setting up the interaction configuration can be found in \citeauthor{Ber99a} (1999\,a,b). For the simulations in this paper we have chosen two different impact times, one at $t_{\rm i}\!=\!60$ and the other at $t_{\rm i}\!=\!150$. As will be shown in the next sections, the former corresponds roughly to the time of the maximum bar strength and the latter to a time when the bar is considerably weaker. We will thus hereafter refer to cases with the first impact time as interactions with a strong bar and to cases with the second one as interactions with a weak bar. \begin{table} \caption{Initial model parameters} \begin{tabular}{cccrlccc} \hline & & Type & N$_D$ & M$_D$ & a$_D$ & R$_{\rm cut}$ & z$_0$ \\ \hline {\bf Disc} & stars & KT & 13\,500 & 0.56 & 1.0 & 5.0 & 0.20 \\ & gas & KT & 10\,000 & 0.14 & 1.0 & 5.0 & 0.05 \\ {\bf Halo} & stars & PL & 32\,500 & 1.30 & 5.0 & 10.0 & $\cdots$\\ {\bf Companion}& stars & PL & 10\,000 & 0.40 & 0.195 & 3.0 & $\cdots$ \\ \hline \end{tabular} \end{table} \section{Morphological Evolution} \subsection{Isolated Galaxy} The model was constructed so as to be globally unstable to non-axisym\-metric perturbations and forms a large-scale bar in a few disc rotations ($t_{\rm rot}$, measured at disc half-mass radius). The bar reaches its maximum strength at $t\!=\!60$ or some $20\,t_{\rm rot}$. At this time the bar has a major axis length of $a\!=\!6$\,kpc and an axial ratio of $\sim\!3\!:\!1$. Both stellar and gaseous trailing spiral arms emerge from the end of the bar. While the stellar arms slowly dissolve and are hardly visible at $t\!=\!100$, large spiral features in the gas persist throughout the run. The gas also forms straight off-set shocks at the leading edges of the stellar bar . Due to the gravitational torque of the bar, the gas is driven towards the center and within two bar rotations some $50$ percent of the total gas mass collects in an oval nuclear disc elongated along the bar. As a result of the growing mass concentration at the center the amplitude of the bar decreases, but then settles down to a quasi-steady state as the initial burst of inflow slows down. \begin{figure} \centerline{\epsfig{file=figure1.ps, width=12cm}} \caption{Isolated model. Shown is the evolution of the gaseous disc only.} \end{figure} The inner regions are connected with the outer disc through two trailing spiral arms, which by way of their shocks feed material inward to the growing central disc. Around the nuclear disc forms another disc which is oriented perpendicular to the bar major axis. This indicates the presence of an ILR, i.e. the existence of an $x_2$ orbit family. Outside the central region there is a deficiency of gas out to a radius of 3\,kpc, where a gas ring forms close to the UHR. A detailed discussion on orbits and resonances in isolated gas-rich barred galaxies can be found in \inlinecite{Ber98}. \subsection{Central Passage} In the model with a strong bar both the stellar and the gaseous components of the bar get torn apart into two fragments. The stellar fragments merge again in about $3\,t_{\rm rot}$. The gaseous fragments do not merge as fast, but flow back and forth inside the bar potential avoiding the center for some $6\,t_{\rm rot}$ and appear as two separate nuclei, until they finally merge to form a dense nuclear disc. A lower density, more extended circumnuclear disc forms around the denser nuclear disc as in the isolated model, but with a much larger radius. \begin{figure} \centerline{\epsfig{file=figure2.ps, width=12cm}} \caption{Central impact with strong bar. Shown is the evolution of the gaseous disc only.} \end{figure} The central impact also creates ring-like expanding density waves in the disc, similarly to what happens for non-barred galaxies (\opencite{LT76}). The first ring reaches as far as half the corotation radius and then fragments, with much of its mass flowing back toward the inner few kpc. A second ring propagating outwards follows the first, with trailing spokes forming between the two. For the weak bar the rings are almost perfectly circular, while for the strong bar they are more asymmetric. One could expect such density enhancements to be accompanied by the formation of bright young blue stars and H{\rm II} regions. The expansion velocity of the first ring drops gradually from 88\,km/s to roughly 35\,km/s. The second ring starts with a lower expansion velocity of 29\,km/s, decreasing further to 12\,km/s, which is the sound speed of the isothermal gas in the model. With the passage of the two rings the stellar disc is heated up, preventing the formation of a third stellar ring. In contrast, more than two rings form in the gas, but from the third onwards they are very weak and dissolve quickly, since they are not supported by an accompanying stellar density enhancement. These rings eventually merge with the existing spiral arm segments. \subsection{Minor Axis Passage} In these simulations the companion hits the disc along the bar minor axis at a distance of approximately $r_{\rm i}\!=\!3.0$\,kpc. In both strong and weak bar models the stellar bar gets displaced from the center immediately following the impact to a distance of approximately 3.0\,kpc in the direction of the impact point. This offset lasts for a period of roughly $\Delta t\!=\! 0.6$\,Gyr. An expanding circular ring is formed at the impact point, which upon encountering the bar and spiral arms becomes distorted. The ring expands more freely on the side of the bar which is opposite the impact point and also maintains longer its circular shape, before merging with a spiral arm. This merged ring/arm then opens up forming a large spiral feature. On its other side, the ring stays close to the bar. Also a new spiral arm segment forms, connected to one end of the still displaced bar. At the time when the bar is again nearly centered a second ring starts to form, expanding outwards from the impact point. This ring merges with the newly formed spiral arm segment. Between this feature and the large spiral arm there is a noticeable deficiency of both stars and gas. Similar to the central impact, additional rings form at later times and merge with the many spiral arms that are present, forming spiral segments which wind up and dissolve over time. The final gas morphology is similar to that of the isolated case, even though the stellar bar has been destroyed. \begin{figure} \centerline{\epsfig{file=figure3.ps, width=12cm}} \caption{Major axis impact with strong bar. Shown is the evolution of the gaseous disc only.} \end{figure} \subsection{Major Axis Passage} In these simulations the companion hits the disc on the bar major axis near corotation. Immediately following the passage of the companion through the disc plane, the stellar bar is displaced from the center to a distance of 2.4\,kpc in the direction of the impact point . In general we find the same kind of evolution as in the minor axis impact, though the lengths and orientations of the generated features differ. The most significant difference is that the bar survives the major axis impact. In fact the evolution of the bar strength follows closely that of the isolated case. Again, the final gas configuration is similar to the other models. \section{Evolution of the bar} \label{bar_evol} \subsection{Amplitude of the bar} We have measured the bar strength as defined by the amplitude of the $m\!=\!2$ Fourier component of the stellar mass distribution in the disc. For models in which the bar gets displaced we trace the center of mass of the stellar bar and measure its amplitude within a radius of 3\,kpc from this center. The results are shown in figure~2\,(a) and (b). \begin{figure}[b] \centerline{\epsfig{file=figure4a.eps, width=4.8cm, angle=-90} \epsfig{file=figure4b.eps, width=4.8cm, angle=-90} } \caption{Evolution of bar strength. Models at impact times of (a) $t=60$ and (b) $t=150$ are shown.} \end{figure} With the central passage the bar is destroyed immediately following the impact of the companion, independent of the bar strength. In the model with the minor axis passage the bar is also destroyed, but its amplitude decreases more slowly over time than in the central passage. Impacts along the bar major axis show the same evolution of the bar strength as in the isolated model. \subsection{Pattern speed of the bar} In the isolated model the pattern speed increases linearly with time during the period in which gas is flowing inward. Once the inflow stops the bar quickly settles down to a quasi-steady state and rotates with the constant rate of $\Omega_p\!=\!0.3\tau^{-1}$. In the case of the central passage, the bar is quickly weakened and it is difficult to follow the pattern speed for more than a few disc rotations after the impact. However, the general trend is that of a slowing down of the bar in both the strong and weak bar models. In the strong bar minor axis passage the pattern speed increases with time, converging to a value of $\Omega_p\!=\!0.5$. In contrast, the weak bar minor axis passage shows little variation of the pattern speed. The major axis passages show the same evolution of pattern speed as in the isolated model. \section{Summary} Using N-body/SPH simulations we have examined the effect of a small spherical companion passing through the disc of a gas-rich barred galaxy. The impact position has been varied with respect to the bar and two different impact times have been chosen, corresponding to weak and strong bar cases. The interaction produces characteristic features like expanding rings, spokes, and an off-centered bar, similar to what has been found in pure N-body simulations. The offset passages lead to a displacement of the bar from the center which lasts for approximately 0.6\,Gyr. While the morphology for several disc rotation times following the impact is drastically altered by the interaction, the overall final state of the bar and disc is similar to that of the isolated case. However, the interaction does change the detailed structure of the gas distribution in the central region and signs of an interaction in the form of asymmetries are left in the outer disc. The bar is not necessarily destroyed by the interaction, even though being temporarily displaced from the center.
\section{Introduction} \label{Sec:Intro} In a fully-interacting theory, the number of constituents of a hadron is not conserved, and the picture of a hadron as being comprised of three quarks (baryon) or a quark and antiquark (meson) is na{\"\i}ve. One way to go beyond a picture in which hadrons are simple bound states of valence quarks is to consider their mixing with multiple-hadron states. The success of models which describe hadrons in terms of only a few degrees of freedom, such as the constituent-quark model, suggests that mixing with multiple-hadron states should represent only a small correction to the predominant valence quark structure of the hadron. Some observables may be sensitive to contributions that arise from mixings of hadrons with multiple-hadron states. These observables would provide a means to probe the structure of hadrons beyond that of the valence quarks. One such observable is the mass splitting between the $\rho$ and $\omega$ mesons, which is the subject of this article. The $\rho^0$ and $\omega$ have a similar quark substructure, differing only in their total isospin (and hence $G$-parity). SU(3)-flavor violating contributions due to the different masses of the $u$, $d$ and $s$ quarks tend to split the masses of the $\rho$ and $\omega$ mesons. However, these contributions alone do not account for most of the mass splitting \cite{MT}. Other contributions arise from the mixing of the $\rho$ and $\omega$ mesons with multiple-hadron states of differing $G$-parity. An important example of this is the shift in the mass of the $\rho$ meson due to its mixing with the two-pion state. As mixing between the $\omega$ meson and the two-pion state is forbidden by $G$-parity, the mass of the $\omega$ meson doesn't receive such a contribution from the two-pion state. The difference between the $\rho$ and $\omega$ mesons due to mixing with all possible multiple-hadron states results in an {\em observable} splitting of the masses $m_{\rho}$ and $m_{\omega}$, which would otherwise be degenerate (in the absence of SU(3)-flavor breaking). In \Ref{LC}, the mass shift of the $\rho$ meson due to its mixing with the two-pion state was investigated using an effective chiral Lagrangian approach. A form factor was introduced into the resulting dispersion relations for the $\rho$-meson self energy $\Pi_{\mu \nu}(q^2)$ to render it finite. Several different forms of this form factor were considered and each produced similar results. For example, using a dipole form factor, it was found, over a large range of dipole widths $0.6 \leq \Lambda^2 \leq $ 4.0~GeV$^2$, that the $\rho$-meson mass is shifted by roughly $-10$ to $-20$ MeV due to its coupling with the two-pion state. The purpose of the study of \Ref{LC} was to determine whether results obtained by lattice studies of QCD for the $\rho$ meson mass are significantly affected when the $\rho$ meson is ``unquenched''; that is, when the $\rho$ meson is allowed to mix with two-pion states. Such mixings with multiple-hadron states are typically neglected in lattice studies. It is clear from the study, which relies on very little dynamical modeling, that the inclusion of intermediate meson states into the vector-meson self energy $\Pi_{\mu \nu}(q^2)$ is indeed a small effect which can be treated as a correction to the predominant quark-antiquark structure of the vector meson. An extensive study of the $\rho$-$\omega$ mass splitting is carried out in \Ref{GI}. In this work the contributions of many different two-meson intermediate states to the masses of the $\rho$ and $\omega$ are evaluated within a nonrelativistic framework, using time-ordered perturbation theory to evaluate the loops. The three-meson vertices and their momentum dependence are evaluated using a string-breaking picture of the strong decays based on the flux-tube model of the confining interaction, and quark pair creation with vacuum ($^3P_0$) quantum numbers. It is noted that the $\rho$-$\omega$ splitting is directly proportional to an OZI-violating mixing between ground-state vector $u\bar{u}$ and $d\bar{d}$ mesons. This mixing is due to two-meson intermediate states which can link the initial and final mesons. It is shown in \Ref{GI} that, in the closure limit where the energy denominators for the intermediate states are constant, the $^3P_0$ pair-creation operator cannot link these two vector states, so that the OZI rule is exact in this limit. The presence of a small OZI-violating mass splitting is, therefore, interpreted as a deviation of the sum over intermediate states from this closure limit due to the differing properties of the physical intermediate states. Although the authors of \Ref{GI} point out that their calculation of this mixing is not formally complete due to the absence of pure annihilation diagrams where the intermediate state contains only glue (which can be written as a sum over heavier glueball states and are likely suppressed), the sum over intermediate states is much more extensive than that carried out here. The convergence of this sum, although convincingly demonstrated, is rather slow in their calculation, requiring many different meson pairs with many possible relative angular momenta. This can be illustrated by examining the size of individual terms in the sum; for example, the $\pi\pi$ loop contributes $-142$ MeV to the $\rho$ mass, and the $\omega\pi$ loop contributes $-146$ MeV (the corresponding $\rho\pi$ loop contributes roughly three times this amount to the $\omega$ mass). It is pointed out in \Ref{GI} that without a flux-tube overlap function and a form factor at the pair creation vertex, both of which are necessary for a theoretically consistent picture of the strong decays and suppress high-momentum created pairs, this sum would not formally converge. Note that neither of these features of the strong decay model are essential to fitting the strong decays of mesons in this picture, but were essential to the convergence of the loop expansion in \Ref{GI}. A global fit to the meson decays was carried out in order to fit the decay model (in particular the pair creation vertex form factor). It was concluded in \Ref{GI} that reasonable results for the $\rho$-$\omega$ mass splitting could be obtained only when the effects of many two-meson intermediate states are included in the loop expansion. Such a conclusion raises questions about the validity of the usual picture of a meson as being well described as a bound state of a valence quark and antiquark. The motivation for carrying out the present study is to attempt to reconcile the small effects on the $\rho$ mass due to the two-pion loop found in \Ref{LC} with the more substantial mass shift found in \Ref{GI}, and to examine the consequences for the convergence of the sum over intermediate states performed in \Ref{GI}. The $\rho$-$\omega$ mass splitting is evaluated using meson transition form factors obtained from a covariant, quantum field theoretic model based on the Schwinger-Dyson equations of QCD. A similar approach was employed in \Ref{MT} for the $\pi\pi$ intermediate state. The present study extends the work of \Refs{MT,LC} by considering the additional channels $K \bar{K}$, $\omega \pi$, $\rho \pi$, $\omega \eta$, $\rho \eta$, and $K^{*}K$. (In \Ref{Hollenberg}, a non-local quark four-point interaction was used to study the $\rho$-$\omega$ mass splitting and it was shown that the three-meson intermediate state $\pi\pi\pi$ produces a negligible mass shift. Hence, three-meson intermediate states are neglected in the present study.) The model employed is fit to various properties and electromagnetic form factors of the $\pi$ and $K$, and has been shown to give reasonable results for the decays and electromagnetic form factors of the light vector mesons in \Ref{Hawes}. Although a very different framework is employed here, results that are qualitatively and quantitatively similar to those found in \Refs{LC} are obtained for the $\rho$-meson mass shift due to the two-pion loop. In each of the seven channels considered here, a mass shift of less than $10\%$ of the total mass of the $\rho$ is observed. These mass shifts partially cancel each other when calculating the difference $m_{\omega} - m_{\rho}$, as observed in \Ref{GI}, and we obtain a net mass splitting of $m_{\omega} - m_{\rho} \approx $ 25~MeV. The experimentally observed value for this mass splitting is $m_{\omega} - m_{\rho}=12\pm 1$~MeV. Of course, the calculation of the $\rho$-$\omega$ mass splitting described herein is incomplete. Other channels, such as the two-vector-meson intermediate states, might be expected to contribute significantly \cite{GI}. Nonetheless, these results are consistent with those obtained in \Ref{LC}, and indicate that the sum over two-meson intermediate states should converge quite rapidly as the masses of the intermediate state mesons increase, in contrast to what is found in \Ref{GI}. The dependence of the vector meson self energy on the masses of the intermediate mesons is also explored in detail. It is shown that contributions from two-meson loops with meson masses above $m_\rho/2$ contribute only small amounts to the mass shift, and their contributions decrease rapidly as the masses of the intermediate mesons are increased. This is the essential mechanism for the expected rapid convergence of the sum over two-meson loops and is insensitive to the details of the model meson transition form factors, the only dynamical input employed in the present calculations. Another observable that is sensitive to the inclusion of meson loops is also explored in this study for the first time. With the same $\rho\rightarrow\pi\pi$ transition form factors used in the calculation of the $\rho$-$\omega$ mass splitting, the contribution of $\pi\pi$ loops to the $\rho$-meson electromagnetic form factor is calculated. This model predicts a $10\%$ increase in the charge radius of the $\rho$ meson relative to that obtained in \Ref{Hawes} which considered only the quark-antiquark structure of the vector meson. The inclusion of pion loops leads to a similar $10$--$15\%$ increase for the $\pi$-meson charge radius \cite{Bender}. The results from the present study also indicate that of the two-meson states considered, the two-pion state provides the most significant correction to the vector meson charge radius. The brief lifetime of the $\rho$ meson makes a direct measurement of its electromagnetic charge radius impossible at present. However, the size of the neutral vector meson can be extracted from high-energy data for diffractive, vector meson photoproduction on a nucleon. It is possible, using a semi-empirical model of diffractive photoproduction \cite{MAP}, to relate the {\em diffractive radii} of mesons to their charge radii. Given the dominance of the two-pion state this implies that the increase in the charge radius of the $\rho$ meson due to the two-pion loop corresponds to an analogous increase of the diffractive radius of the $\rho$ meson. Since $G$-parity forbids the $\omega$ meson from receiving such contributions from the two-pion state, a measured difference in the diffractive radii of the $\rho$ and $\omega$ mesons can provide a means to probe the magnitude of the two-pion contribution to $\rho$ meson observables. The organization of this article is as follows. In \Sec{Sec:Amps}, the general Lorentz and flavor structure of the meson transition amplitudes, necessary for the calculation of the vector-meson self energy and electromagnetic form factors, are derived. These elements are then employed in a study of the Schwinger-Dyson equation for the vector meson propagator in \Sec{Sec:SelfEnergy}. The mass of the vector meson is related to the real part of the vector-meson self energy, and the total decay width of the vector meson is related to the imaginary part. These same elements are then employed in a study of the $\rho$ meson electromagnetic form factors in \Sec{Sec:EMFF}. In \Sec{Sec:Orders}, the contributions to the vector-meson self energy arising for the different time orderings of the intermediate meson propagators are investigated. Aspects of time-ordered quantum mechanical frameworks, such as that employed in \Ref{GI}, and the Lorentz-covariant Euclidean-based quantum field theoretic framework employed here are discussed and contrasted. In \Sec{Sec:Model}, the dynamical model used to calculate the off-mass-shell behavior of the meson transition amplitudes is described. The model describes the meson transitions in terms of loops involving nonperturbatively-dressed quarks and model Bethe-Salpeter amplitudes that describe the quark-antiquark substructure of the mesons. The dressed-quark propagators and Bethe-Salpeter amplitudes employed herein were developed and tested in numerous studies of meson observables \cite{Hawes,Burden,NewOnes} and in numerical studies of the quark Schwinger-Dyson equation of QCD \cite{Frank,Maris,Wightman}. Results for the $\rho$- and $\omega$-meson self energies are given in \Secs{Sec:ResultsMassDep} and \ref{Sec:ResultsMassShifts}. In \Sec{Sec:ResultsOrders}, numerical results for each of the possible time orderings of the two-pion-loop contribution to the self energy are provided. Results for $\rho$-meson electromagnetic form factor $G_E$ and charge radius are given in \Sec{Sec:ResultsEMffs}. The conclusions of this study are given in \Sec{Sec:Concl}. \section{Meson-loop contributions} \label{Sec:Method} In this section some of the formalism necessary to describe the effects of two-meson loops on the vector-meson self energy and electromagnetic form factors is presented. In \Sec{Sec:Amps}, the most general Lorentz-, parity- and SU(3)-flavor-covariant amplitudes that describe the coupling of a vector meson to two pseudoscalar mesons or a vector and a pseudoscalar meson are presented. The transition amplitudes are written in terms of a coupling constant and an {\em off-mass-shell} transition form factor. Both are defined so that the transition form factor is equal to one when all three mesons are on shell. In the case of the $\rho$ meson coupling to two pions, the coupling constant associated with the transition amplitude is related to the decay $\rho \rightarrow \pi \pi$ and is found by experiment to be $g_{\rho \pi \pi}$ = 6.03. The other coupling constants considered herein as well as the off-mass-shell transition form factors are obtained from considerations of SU(3)-flavor invariance or the dynamical model described in \Sec{Sec:Model}. In \Sec{Sec:SelfEnergy}, the Schwinger-Dyson equation for the vector meson propagator is given in terms of the vector meson self energy $\Pi_{\mu \nu}(q)$. Expressions are derived that give the self energy in terms of loop integrations involving the propagators for the two intermediate mesons and the transition amplitudes described in \Sec{Sec:Amps}. The imaginary part of the vector-meson self energy is related to the total decay width of the vector meson, and the real part of the self energy provides a contribution to the vector meson mass. In \Sec{Sec:EMFF}, the contribution of the two-pion loop to the $\rho$-meson electromagnetic form factors is considered. A correction to the electromagnetic (EM) charge radius of the $\rho$ meson due to pion loops is obtained, and is compared to the results obtained from a study of the quark-antiquark contribution from \Ref{Hawes}. It is argued that since $G$-parity forbids the $\omega$ meson coupling to two pions, and since the two-pion contribution dominates, measurement of the difference of the $\rho$ meson and $\omega$ meson charge radii provides a means to determine the effects of including such intermediate meson loops. The electromagnetic form factors are given in terms of the same meson transition amplitudes used in \Sec{Sec:SelfEnergy} to calculate the self energies and decay width of the $\rho$ meson. Hence, the calculation of the contribution of pion loops to the $\rho$ meson EM form factor provides a further test of the self-consistency of the present approach. \subsection{Meson transition amplitudes} \label{Sec:Amps} The most general coupling of a $\rho$ meson to two pseudoscalar mesons ($\pi$ or $K$) can be written in terms of the following action: \begin{equation} S \!=\!\! \int\!\! d^4\!x\, d^4\!y\, d^4\!z\,\pi^i(x) \pi^j(y) \rho_{\mu}^k(z) \Lambda^{ijk}_{\mu}(x\-z,y\-z), \end{equation} where $\rho^k_\mu(z)$ is the $\rho$ meson field with flavor index $k$ and Lorentz index $\mu$, and $\pi^i(x)$ is the pseudoscalar meson field with flavor $i$. The nonlocal three-point coupling is described by the amplitude \begin{eqnarray} \Lambda^{ijk}_{\mu}(x\-z,y\-z) & = & \int\!{d^4p_1\over (2\pi)^4}{d^4p_2\over (2\pi)^4} \nonumber\\ & & e^{-ip_1(x-z)-ip_2(y-z)} \Lambda^{ijk}_{\mu}(p_1,p_2) , \end{eqnarray} which can be written in terms of a coupling constant $g_{ijk}$, where $i$, $j$, and $k$ are SU(3)-flavor indices, and a form factor $f^{VPP}(p_1,p_2)$ which is a function of the Lorentz invariants $p_1^2$, $p_2^2$ and $q^2=(p_1+p_2)^2$. The form factor $f^{VPP}(p_1,p_2)$ is defined to be equal to unity, when all three mesons are on their mass shell; that is, $f^{VPP}(p_1,p_2) = 1$ when $p_1^2 = p_2^2 = - m_\pi^2$, $q^2 = (p_1+p_2)^2 = - m_\rho^2$. Hence, one may write: \begin{equation} \Lambda^{ijk}_{\mu}(p_1,p_2) = \frac{1}{2}(p_1\-p_2)_{\mu} \; g^{ijk} \; f^{VPP}(p_1,p_2). \label{VPPvertex} \end{equation} In general, \Eq{VPPvertex} may also have a term which is proportional to $(p_1+p_2)_{\mu}$. Even if such a term were present, it could not contribute to the self energy of an {\em on-shell} vector meson since its contraction with a spin-1 polarization vector is zero. For $\rho \rightarrow \pi \pi$, $g^{ijk} \equiv g_{\rho\pi\pi} \epsilon^{ijk}$. From the above amplitude, one obtains the invariant Feynman amplitude for the decay of, say, the $\rho^0$ into two charged pions: \begin{eqnarray} \lefteqn{ \< \pi^+(p_1); \pi^-(p_2) | T | \rho^0(q,\lambda) \> } \nonumber \\ &=& \varepsilon_{\mu}(q,\lambda) \left[ \Lambda^{-+3}_{\mu}(p_1,p_2) + \Lambda^{+-3}_{\mu}(p_2,p_1) \right], \\ &=& 2 g_{\rho\pi\pi} \; p_1 \cdot \varepsilon(q,\lambda) , \end{eqnarray} where $\varepsilon_{\mu}(q,\lambda)$ is the polarization vector for a $\rho$ meson of momentum $q$ and helicity $\lambda$. The resulting $\rho\rightarrow\pi\pi$ decay width is \begin{equation} \Gamma_{\rho\rightarrow\pi\pi}=\frac{g^2_{\rho\pi\pi}}{4\pi} \frac{m_\rho}{12} \left[1-\frac{4m^2_\pi}{m^2_\rho}\right]^{3/2} . \label{VPP:DecayWidth} \end{equation} The meson transition form factor $f^{VPP}(p_1,p_2)$ does not appear in the relation for the decay width because it is unity when all mesons are on their mass shell. Its value away from this point is therefore not directly observable and must be calculated from a dynamical model. Similarly, the coupling of an $\omega$ meson to a $\rho$ and $\pi$ meson is described by the action: \begin{equation} S \!= \! \int\!\! d^4\!x\, d^4\!y\, d^4\!z\, \rho_{\mu}^i(x) \pi^j(y) \omega_{\nu}(z) \Lambda_{\mu\nu}^{ij}(x\-z,y\-z) . \end{equation} Here, $\omega_{\nu}(z)$ is the $\omega$-meson field and $\Lambda_{\mu\nu}^{ij}(x\-z,y\-z)$ is the nonlocal transition amplitude describing the coupling of two vector mesons to a pseudoscalar meson. Introducing the coupling constant $g_{\omega\rho\pi}$ and form factor $f^{VVP}(p_1,p_2)$, considerations of Lorentz covariance give the Fourier transform of $\Lambda_{\mu\nu}^{ij}(x\-z,y\-z)$ as \begin{equation} \Lambda_{\mu\nu}^{ij}(p_1,p_2) = \epsilon_{\mu\nu\alpha\beta}{p_{1\alpha}p_{2\beta} \over m_{\rho}} \; \delta^{ij} g_{\omega\rho\pi} f^{VVP}(p_1,p_2) \label{VVPvertex} \end{equation} where $p_1$ and $p_2$ are the momenta of the $\rho$ and $\pi$ mesons, respectively, and $f^{VVP}(p_1,p_2) = 1$ when $q^2 = (p_1+p_2)^2 = -m^2_\omega$, $p_1^2 = -m^2_\rho$, and $p_2^2= -m^2_\pi$. Unlike the previous case, a lack of phase space prevents the decay $\omega \rightarrow \rho \pi$, so that it is impossible to determine the value of the coupling constant $g_{\omega\rho\pi}$ from experiment. Hence, both the meson-transition form factor $f^{VVP}(p_1,p_2)$ and the coupling constant $g_{\omega\rho\pi}$ must be determined from a model calculation. In \Sec{Sec:Model}, a dynamical model is described and used to determine the transition form factors $f^{VPP}(p_1,p_2)$ and $f^{VVP}(p_1,p_2)$, and the coupling constant $g_{\omega\rho\pi}$. The model has been shown to give good results for the similar, physically accessible decays $\rho\to \pi\gamma$ and $K^*\to K\gamma$, in which a vector meson decays into a vector and pseudoscalar \cite{Hawes}. The value of the coupling constant $g_{\rho\pi\pi}$ = 6.03 used herein is extracted from the experimentally observable $\rho\to \pi\pi$ decay width using \Eq{VPP:DecayWidth}. \subsection{Vector-meson self energy} \label{Sec:SelfEnergy} The spin-1 nature of a vector meson entails the following form for the vector-meson self energy when the vector meson is on mass shell ($q^2 = -m_{V}^{2}$): \begin{equation} \Pi_{\mu\nu}(q)=\left(\delta_{\mu\nu}-{q_\mu q_\nu\over q^2}\right)\Pi(q^2) \label{Pimunu} \end{equation} Away from the on-mass-shell point $q^2 = - m_{V}^2$, the vector-meson self energy is of the general form: \begin{equation} \Pi_{\mu\nu}(q)=\left(\delta_{\mu\nu}-{q_\mu q_\nu\over q^2}\right)\Pi(q^2) + \frac{q_{\mu} q_{\nu}}{q^2} \Theta(q^2). \label{GenPimunu} \end{equation} To project out the Lorentz scalar function $\Pi(q^2)$, a projection operator $\sfrac{1}{3} {\bf T}_{\mu \nu}(q,q^2)$ where \begin{equation} {\bf T}_{\mu\nu}(q,-m^2) \equiv \delta_{\mu \nu} + q_{\mu} q_{\nu} / m^2 \label{Tdef} . \end{equation} is employed. In the following, the dressed vector meson propagator $\tilde{D}_{\mu\nu}(q)={\bf T}_{\mu\nu}(q,-m_V^2) \tilde{D}(q^2)$ is given in terms of the ``bare'' vector meson propagator $D_{\mu\nu}(q) = {\bf T}_{\mu\nu}(q,-m_V^2) {D}_0(q^2)$ and the vector-meson self energy $\Pi_{\mu\nu}(q) = {\bf T}_{\mu\nu}(q,-m^2_V) {\Pi}(q^2)$ which describes the mixing of the vector meson with meson loops (and in general, multiple-hadron intermediate states). The ``bare'' vector-meson propagator $D_0(q^2)$ is presumed to arise from the quark-antiquark structure of the vector meson, and does not include the effects of multiple-meson intermediate states. Defined in this manner, quark confinement requires that the ``bare'' vector-meson propagator $D_0(q^2)$ is a real-valued function of $q^2$. The presence of a non-zero imaginary part for $D_0(q^2)$ would entail a loss of flux for vector-meson propagation, which is a signal for the existence of open decay channels for the vector meson. In our approach, such decay channels arise from the coupling of the vector meson with multiple-hadron states. Mixing the vector meson with multiple-hadron states results in a dressed, vector-meson propagator $\tilde{D}(q^2)$ that is a complex-valued function of the vector-meson momentum $q^2$. The dressed, vector-meson propagator $\tilde{D}(q^2)$ is obtained by solving the Schwinger-Dyson equation: \begin{equation} \tilde{D}=D_0 + D_0 \Pi \tilde{D}. \label{DSE} \end{equation} This integral equation gives the dressed vector meson propagator $\tilde{D}(q^2)$ in terms of the bare propagator $D_0(q^2)$ by dressing the vector meson by two-meson loops as given by $\Pi(q^2)$. As an integral equation, \Eq{DSE} necessarily implies the insertion of infinitely many self-energy $\Pi(q^2)$ contributions into the dressed, vector-meson propagator $\tilde{D}(q^2)$; hence, the dressing of the vector-meson propagator resulting from \Eq{DSE} is inherently nonperturbative. The solution to the Schwinger-Dyson equation is \begin{eqnarray} \tilde{D}&=&D_0+D_0 \Pi D_0 + D_0 \Pi D_0 \Pi D_0 + \cdots , \nonumber\\ && =D_0 \left( \frac{1}{1 + \Pi D_0} \right), \end{eqnarray} so that \begin{equation} \tilde{D}^{-1}=(1 + \Pi D_0)D_0^{-1}=D_0^{-1} + \Pi . \label{Dinv} \end{equation} Near $q^2 = - m_{V}^2$, the difference between the dressed and bare inverse propagators is given by \begin{eqnarray} \lefteqn{ \Pi(q^2) = \tilde{D}^{-1}(q^2) - D_0^{-1}(q^2) } \nonumber \\ &=& (1 + \delta Z) \; (q^2+m_V^2 - i m_V \Gamma) \label{DefZ} -(q^2+m_0^2), \end{eqnarray} where $m_0$ is the bare mass of the vector meson, $\Gamma$ is the total decay width of the vector meson, and $1 + \delta Z$ is the field renormalization for the vector meson, chosen to ensure that the asymptotic vector meson has unit normalization. In \Sec{Sec:Model}, the Bethe-Salpeter amplitudes, which describe the quark-antiquark substructure of the meson, are normalized to unity. Thus, from the definition of the normalization $1 + \delta Z$, as given by \Eq{DefZ}, it follows that the {\em correction} to the normalization that results from the inclusion of meson-loops is given by $\delta Z$. For meson loops to be considered as a small effect, $\delta Z$ must be smaller than one. Comparison of \Eq{DefZ} and \Eq{Dinv} gives \begin{eqnarray} -{\rm Im}(\Pi)&=& (1 + \delta Z) \, m_V \,\Gamma \\ {\rm Re}(\Pi)&=& \delta Z q^2 + (1 + \delta Z) m_V^2 - m_0^2 \label{RePi} . \end{eqnarray} \Eq{RePi} provides a means to determine the normalization $\delta Z$ from the self energy $\Pi(q^2)$: \begin{equation} {d\over dq^2}\left[{\rm Re}\Pi(q^2)\right]_{q^2=-m_V^2}=\delta Z, \label{Zdef} \end{equation} from which it follows: \begin{equation} m_V^2=m_0^2+{\rm Re}\Pi(-m_V^2) . \end{equation} This equation gives the mass of the vector meson $m_V$, in terms of the bare mass $m_0$ (arising from the quark-antiquark substructure of the vector meson) and the self energy $\Pi(q^2)$. It can be solved iteratively, \begin{eqnarray} m_V^2&=&m_0^2+{\rm Re}\Pi\left(-m_0^2-{\rm Re}\Pi(-m_V^2)\right) \nonumber\\ &\approx&m_0^2+{\rm Re}\Pi(-m_0^2) -{\rm Re}\Pi^\prime(-m_0^2){\rm Re}\Pi(-m_0^2), \label{MassShift} \end{eqnarray} where the expansion assumes small $m_V^2-m_0^2$. This relation yields a first-order correction to the vector meson mass: \begin{equation} m_V^2 \approx m_0^2+{\rm Re}\Pi(-m_0^2). \label{massShifts} \end{equation} From \Eq{Zdef}, one identifies ${\rm Re}\Pi^\prime(-m_V^2)$ with $\delta Z$. Thus, the second order correction to the vector meson mass shift is $- \delta Z\,{\rm Re}\Pi(-m_0^2)$, which is suppressed by a factor $\delta Z$ compared to the first-order term in \Eq{MassShift}. The importance of higher-order corrections can be estimated from the momentum dependence of $\Pi(q^2)$ in the vicinity of $m_0^2$. It is found to be sufficient (see \Sec{Sec:Results}) to include only the lowest-order term from \Eq{MassShift} to obtain a good estimate for the dressed, vector-meson mass $m_V$. As stated above, the self energy $\Pi(q^2)$ describes the dressing of the vector meson due to two-meson loops. It can be described in terms of a fluctuation of the vector meson into a multiple-meson intermediate state. This is given in terms of meson transition amplitudes, the form of which was described in \Sec{Sec:Amps}, and propagators for the intermediate mesons. The contribution to the vector-meson self energy $\Pi(q^2)$ due to a two-pseudoscalar-meson loop is then given by \begin{eqnarray} \Pi^{kk^\prime}_{\mu\nu}(q) &=& \-2\! \int \!\!\frac{d^4\!k}{(2\pi)^4} \Lambda_{\mu}^{ijk}(\-p_1,\-p_2) \Delta(p_1,m_P) \nonumber \\ & & \times \Delta(p_2,m_{P'}) \Lambda_{\nu}^{ijk^\prime }(p_1,p_2) \nonumber \\ &\equiv & \delta^{kk^\prime}\Pi_{\mu\nu}(q), \label{PimunuPP} \end{eqnarray} where $m_P$ and $m_{P'}$ are the masses of the pseudoscalar mesons, $p_1 = \sfrac{1}{2} q + k$, $p_2 = \sfrac{1}{2} q - k$, and the meson transition amplitude $\Lambda^{ijk}_{\mu}(p_1,p_2)$, given by \Eq{VPPvertex}, describes the transition $V \rightarrow P P$ from a vector meson $V$ to two-pseudoscalar mesons $P$ and $P$ with momenta $p_1$ and $p_2$, respectively. The propagation of the intermediate pseudoscalar mesons is given in terms of the following scalar-boson propagator: \begin{equation} \Delta(p_1,m_P)=(p_1^2+m_P^2-i\epsilon)^{-1}, \end{equation} where $\epsilon$ is an infinitesimal positive number included to ensure the proper boundary conditions for the decay $V \rightarrow P P$. This process is depicted in \Fig{Fig:Loop}, in which the meson transition amplitude $\Lambda_{\mu}^{ijk}(p_1,p_2)$ is given in terms of a quark-loop triangle diagram. In \Sec{Sec:Model}, the transition form factor is calculated using the dressed-quark propagators and Bethe-Salpeter amplitudes of a quantum field theoretic model based on studies of the Schwinger-Dyson equations of QCD. In impulse approximation, the leading contribution to the meson transition amplitude $ \Lambda_{\mu}^{ijk}(p_1,p_2)$ is given in terms of the triangle diagram depicted in \Fig{Fig:Loop}. Applying the projection operator $\sfrac{1}{3}T_{\mu\nu}(q,q^2)$ of \Eq{Tdef} to $\Pi_{\mu\nu}(q)$ in \Eq{PimunuPP} yields the Lorentz-scalar vector meson self energy due to two-pseudoscalar-meson loops \begin{eqnarray} \Pi^{PP}&&(q^2,m_P,m_{P'}) = \nonumber\\ &&-\frac{4}{3} \int\! \frac{d^4\!k}{(2\pi)^4} \Delta (p_1,m_P) \Delta (p_2,m_{P'}) {\cal F}^{PP}(k,q), \label{PiPP} \end{eqnarray} with \begin{equation} {\cal F}^{PP}(k,q)=\left(k^2- \frac{k\cdot q^2}{q^2} \right) \big[g_{\rho\pi\pi} f^{VPP}(p_1,p_2)\big]^2 . \label{FPP} \end{equation} Similarly, the contribution of a vector-pseudoscalar-meson loop to the vector-meson self energy is given by \begin{eqnarray} \Pi^{ii^\prime}_{\mu\nu}(q) &=& \int\! \frac{d^4\!k}{(2\pi)^4} \Lambda_{\mu\sigma}^{ij}(\-p_1,\-p_2) D_{\sigma\tau}(p_1,m_V) \nonumber\ \\ & & \times \Delta(p_2,m_P) \Lambda_{\nu\tau}^{i^\prime j }(p_1,p_2), \label{PimunuVP} \end{eqnarray} where $m_V$ and $m_P$ are the masses of the intermediate vector and pseudoscalar mesons, respectively, and the meson transition amplitude $\Lambda_{\mu \nu}^{ij}(p_1,p_2)$ is given by \Eq{VVPvertex}. The propagator for the intermediate vector meson may be written as \begin{equation} D_{\mu \nu}(p,m_V)=\left( \delta_{\mu \nu} + \frac{p_{\mu}p_{\nu}}{m_V^2}\right) \Delta(p,m_V) . \end{equation} Substituting this and the relation for $\Lambda_{\mu \nu}^{ij}(p_1,p_2)$ from \Eq{VVPvertex} into \Eq{PimunuVP}, and applying the projection operator $\sfrac{1}{3}{\bf T}_{\mu\nu}(q,q^2)$ to the result, yields \begin{eqnarray} \Pi^{VP}&&(q^2,m_V,m_P)=\nonumber\\ &&\frac{2}{3} \int\! \frac{d^4\!k}{(2\pi)^4} \Delta (p_1,m_V) \Delta (p_2,m_P) {\cal F}^{VP}(k,q), \label{PiVP} \end{eqnarray} with \begin{equation} {\cal F}^{VP}(k,q)= \left[\frac{(k\cdot q)^2 - k^2 q^2 }{m_{\rho}^2}\right] \big[ g_{\omega\rho\pi} f^{VVP}(p_1,p_2)\big]^2. \label{FVP} \end{equation} \begin{figure}[t] \centering{\ \mbox{\ \epsfig{figure=fig1.eps,width=2.9cm,angle=-90}}} \vspace*{0.15cm} \caption{ Feynman diagram depicting the quark substructure of the vector-two-pseudoscalar meson transition amplitude $\Lambda_{\mu}^{ijk}(p_1,p_2)$ (shown as a triangle-shaped quark-loop diagram) given by \Eq{LamVPP:Quark} and the intermediate pion propagators (dashed curves) necessary for the calculation of the vector-meson self energy $\Pi^{PP}(q^2,m_P,m_P)$. } \label{Fig:Loop} \end{figure} The flavor structure of the amplitudes in \Eqs{PimunuPP} and (\ref{PimunuVP}) are given for the particular case of the $\rho$ meson self energy. It is easy to generalize the relations in \Eqs{PiPP} and (\ref{PiVP}) to obtain the isoscalar $\omega$-meson self energy. Assuming that the coupling constants for $V\rightarrow P P$ and $V \rightarrow V P$ are SU(3)-flavor symmetric, one can write the self energy of the $\rho$ meson and the $\omega$ meson in terms of the Lorentz-scalar functions $\Pi^{PP}(q^2,m_P,m_{P'})$ and $\Pi^{VP}(q^2,m_V,m_P)$ given by \Eqs{PiPP} and (\ref{PiVP}), respectively, for several possible channels involving the SU(3)-octet of vector and pseudoscalar mesons. For simplicity, only those two-meson loops with the lowest mass intermediate mesons are included in the calculation of the vector-meson self energy $\Pi(q^2)$. In \Sec{Sec:Results}, it is shown that meson loops involving mesons with higher masses are suppressed relative to those with lower masses and from \Ref{Hollenberg}, contributions from three-meson loops are expected to be very small. For the present study, the following channels are included: $\pi\pi$, $K\bar{K}$, $\rho\pi$, $\omega\pi$, $\rho\eta$, $\omega\eta$, and $K^* K$. Hence, the self energy of the $\rho$-meson is given by \begin{eqnarray} \Pi(q^2)&=& \Pi^{PP}(q^2,m_\pi,m_\pi) + \sfrac{1}{2} \Pi^{PP}(q^2,m_K,m_K) \nonumber \\ & & + \Pi^{VP}(q^2,m_\omega,m_\pi) + \Pi^{VP}(q^2,m_{K^*},m_K) \nonumber \\ & & + \Pi^{VP}(q^2,m_\rho,m_\eta) , \label{PiSum:Rho} \end{eqnarray} and that of the $\omega$ meson is given by \begin{eqnarray} \Pi(q^2)&=& 0 + \sfrac{1}{2} \Pi^{PP}(q^2,m_K,m_K) \nonumber \\ && + 3 \Pi^{VP}(q^2,m_\rho,m_\pi) + \Pi^{VP}(q^2,m_{K^*},m_{K}) \nonumber \\ && + \Pi^{VP}(q^2,m_\omega,m_\eta) . \label{PiSum:Omega} \end{eqnarray} Once the meson transition form factors have been determined from a dynamical model, the contributions of the self energies can be obtained from \Eqs{PiSum:Rho} and (\ref{PiSum:Omega}) in a straight-forward manner. As an example, consider the calculation of the loop integral necessary to obtain $\Pi^{PP}(q^2)$, which is illustrated in \Fig{Fig:Loop} and is given by \Eqs{PiPP} and (\ref{FPP}), \begin{eqnarray} \lefteqn{ \Pi^{PP}(q^2,m_P,m_{P'})= -\frac{4}{3} \int\! \frac{d^4\! k }{(2\pi)^4} } \nonumber \\ & & \frac{{\cal F}^{PP}(k,q)} {[(k+q/2)^2+m_P^2-i\epsilon] [(k-q/2)^2+m_{P'}^2-i\epsilon]}. \label{PiPP2} \end{eqnarray} Introducing a Feynman parametrization \begin{equation} \frac{1}{a_1a_2}=\int_0^1 \frac{dx}{[(a_1-a_2)x+a_2]^2} , \end{equation} the denominators of the propagators in \Eq{PiPP2} can be combined. With a change in variables $ k\rightarrow k+\alpha q$, where $\alpha=x-1/2$, two of the integrations can be performed trivially, giving \begin{eqnarray} \lefteqn{ \Pi^{PP}(q^2,m_P,m_{P^\prime}) \hspace*{\fill} }\nonumber \\ &=& \!\- \frac{4}{3} \int_{\-\sfrac{1}{2}}^{\sfrac{1}{2}} \!\! d\alpha \!\!\int \!\! \frac{d^4\! k}{(2\pi)^4} \frac{{\cal F}^{PP}(k \- \alpha q,q)} {[k^{2}\-(\alpha^2\-\sfrac{1}{4}) q^2 \+ \bar{m}^2 \+ \alpha \tilde{m}^2 \- i\epsilon]^2} , \nonumber \\ &=& \- \frac{4}{3} \frac{1}{(2\pi)^3} \int_{\-\sfrac{1}{2}}^{\sfrac{1}{2}} \!\! d\alpha \!\!\int_{\-1}^1 \!\!\! dz \sqrt{1\-z^2} \nonumber \\ & & \times \!\!\int_0^{\infty} \!\! {k^2 \, dk^2} \frac{{\cal F}^{PP}(k\-\alpha q,q)} {[k^2\-(\alpha^2\-\sfrac{1}{4})q^2 \+ \bar{m}^2 \+ \alpha \tilde{m}^2 \-i\epsilon]^2}, \label{loopint} \end{eqnarray} where $z$ is the cosine of the angle between $k$ and $q$, defined by $k\cdot q =\sqrt{k^2}\sqrt{q^2}\,z$, $\bar{m}^2 = \sfrac{1}{2}(m_{P}^2 + m_{P'}^2)$ and $\tilde{m}^2 = (m_{P}^2 - m_{P'}^2)$. An inspection of \Eq{loopint} reveals that the denominator exhibits a zero in the limit that $\epsilon \rightarrow 0$ for $k^2$ equal to \begin{equation} k^2_0 = (\alpha^2 \- \sfrac{1}{4}) q^2 - \bar{m}^2 - \alpha \tilde{m}^2 . \end{equation} This singular behavior leads to a simple pole in the vector-meson self energy that is associated with the decay of the vector meson into two pseudoscalars. The integral in \Eq{loopint} is evaluated numerically by first eliminating one factor of the singularity using \begin{equation} \left(\frac{1}{k^2 - k_0^2}\right)^{2} = \frac{d\;}{dk^2} \; \frac{-1}{k^2 - k_0^2} , \end{equation} integrating $k^2$ by parts twice, and noting that Cauchy's theorem entails \begin{equation} \lim_{\epsilon \rightarrow 0^+} \frac{1}{k^2 - k^2_0 - i\epsilon} = \frac{{\cal P}}{k^2 - k_0^2} + i \pi\delta(k^2 - k_0^2) , \end{equation} where ${\cal P}$ denotes the principal part of the integration. The imaginary part of \Eq{loopint} is obtained quite easily. To evaluate the principal part, the $k^2$-integration domain is divided into two regions. The first region is from the origin ($k^2 =0$) to twice the distance from the pole to the origin ($2 \; k_0^2 $). The second region is from $k^2 = 2 \; k_0^2$ to $k^2 = \infty$. The integrations are numerically calculated using the method of quadratures. In the first region of the integration over $k^2$, each mesh point of the integration is paired with one that is equidistant from the pole at $k^2_0$ to ensure the best cancellation of round-off errors. In the second region, the integration is straight-forward. In \Sec{Sec:Amps}, it is shown that the meson transition amplitudes $f^{VPP}(p_1,p_2)$ decrease rapidly with increasing $k^2$, providing sufficient damping to the integrations in \Eqs{PiPP} and (\ref{PiVP}) to ensure their convergence. Similar techniques are employed in the calculation of the loop integral $\Pi^{VP}(q^2,m_V,m_P)$. \subsection{Vector meson electromagnetic form factor} \label{Sec:EMFF} The emission or absorption of a photon by an on-mass-shell $\rho$ meson is described by the action: \begin{eqnarray} S = \int \!d^4\!x \, d^4\!y \, d^4\!z \; & &\; \rho^i_{\alpha}(x) \; \rho^j_{\beta}(y) \; A_{\mu}(z) \nonumber \\ & & \times \; \epsilon^{ij3} \; \Lambda_{\mu \alpha \beta}(x\!-z,y\!-z) . \end{eqnarray} Charge conjugation and Lorentz covariance constrain the form of the amplitude $\Lambda_{\mu \alpha \beta}(x\!-z,y\!-z)$. Its Fourier transform is written in terms of three Lorentz-invariant form factors $G_i(Q^2)$ for $i=1,2,3$, which depend only on the square of the photon momentum $Q^2 = (-q-q')^2$, \begin{equation} \Lambda_{\mu \alpha \beta}(q,q') = \sum_{i=1}^{3} G_{i}(Q^2) \; T^{[i]}_{\mu\alpha\beta}(q,q') \label{LamSum} , \end{equation} where \begin{eqnarray} T^{[1]}_{\mu \alpha \beta}(q,q^{\prime}) & = & (q_\mu \- q^{\prime}_{\mu}){\bf T}_{\alpha \gamma}(q,-m^2_V) {\bf T}_{\gamma \beta}(q^{\prime},-m^2_V) \;,\nonumber \\ T^{[2]}_{\mu \alpha \beta}(q,q^{\prime}) & = & {\bf T}_{\mu \alpha}(q,-m^2_V) {\bf T}_{\beta \delta}(q^{\prime},-m^2_V) q_\delta \nonumber \\ & & \- {\bf T}_{\mu \beta}(q^{\prime},-m^2_V) {\bf T}_{\alpha \delta}(q,-m^2_V) q^{\prime}_\delta \:, \label{TiDef} \\ T^{[3]}_{\mu \alpha \beta}(q,q^{\prime}) & = & \frac{q_\mu \- q^{\prime}_\mu }{2 m_V^2} {\bf T}_{\alpha \gamma}(q,-m^2_V) q^{\prime}_\gamma {\bf T}_{\beta \delta}(q^{\prime},-m^2_V) q_\delta \;, \nonumber \end{eqnarray} $m_V$ is the mass of the vector meson, and \begin{equation} {\bf T}_{\alpha \beta}(q,-m^2_V) = \delta_{\alpha \beta} + \frac{q_{\alpha} q_{\beta}}{m_V^2} . \end{equation} One can construct a set of tensors $O^{[i]}_{\mu \alpha \beta}$ which can be used to project out from $\Lambda_{\mu\alpha\beta}(q,q^{\prime})$ each of the form factors in \Eq{LamSum}; i.e., $G_i(Q^2)=O^{[i]}_{\mu\alpha\beta}\Lambda_{\mu\alpha\beta}(q,q^{\prime})$. Alternatively, one may construct from linear combinations of the tensors in \Eqs{TiDef} a set of tensors that transform irreducibly under spatial rotations in a frame where there is no energy transferred to the final $\rho$ meson, so $Q_\mu = (\vec{Q},0)$. The form factors associated with this set of tensors are the electric monopole $G_E(Q^2)$, magnetic dipole $G_M(Q^2)$, and electric quadrupole $G_Q(Q^2)$ form factors. They are related to the original set of form factors by \begin{eqnarray} G_E(Q^2) & = & G_1(Q^2) + \frac{2}{3} \; \frac{Q^2}{4 m_V^2} G_Q(Q^2)\:, \nonumber \\ G_M(Q^2) & = & - G_2(Q^2)\:, \label{FFRel} \\ G_Q(Q^2) & = & G_1(Q^2) + G_2(Q^2) + (1 + \frac{Q^2}{4 m_V^2}) G_3(Q^2)\:. \nonumber \end{eqnarray} Similarly, one can construct three tensors that project out these form factors from the transition amplitude in an obvious manner, so that $G_E(Q^2)=O^{[E]}_{\mu\alpha\beta}\Lambda_{\mu\alpha\beta}(q,q^{\prime})$. The amplitude $\Lambda_{\mu \alpha \beta}(q,q')$ receives contributions from many different sources. The electromagnetic properties of the vector meson arise from both its quark substructure as well as from its mixings with other hadron states, such as two-pion intermediate states. The quark-antiquark contribution to the vector-meson EM form factors was explored in \Ref{Hawes} using the same dynamical model described in \Sec{Sec:Model}. The EM form factors that arise from the {\em quark-antiquark} substructure of the vector meson are denoted herein as $G^{q\bar{q}}_{i}(Q^2)$. This present study extends the work of \Ref{Hawes} by exploring the extent to which $\pi$-meson loops contribute to the $\rho$-meson EM form factors, denoted $G_{i}^{\pi\pi}(Q^2)$, and $\rho$-meson charge radius. \begin{figure}[t] \centering{\mbox{\epsfig{figure=fig2.eps,width=8.0cm}}} \caption{ Feynman diagram depicting the impulse approximate two-pion loop contribution to the $\rho$-meson electromagnetic transition amplitude $\Lambda_{\mu\alpha\beta}(p,p')$ of \Eq{Lam2}. } \vspace{1.5ex} \label{Fig:EMpipi} \end{figure} In impulse approximation, the two-pion loop contribution to the $\rho$-meson EM transition amplitude is illustrated in \Fig{Fig:EMpipi} and is given by \begin{eqnarray} \lefteqn{ \Lambda_{\mu \alpha \beta}(q,q') = -4i \!\! \int \!\! \frac{d^4\!k}{(2\pi)^4} \Delta(k_{+}) \Gamma_{\mu}(k_+\+\sfrac{1}{2}Q;Q) } \nonumber \\ & & \times \Delta(k_{+}\+Q) \Lambda^{3-+}_{\beta}(k_{+}\+Q,k_{-}) \Delta(k_{-}) \Lambda^{3+-}_{\alpha}(k_{-},k_{+}) ,\label{Lam2} \end{eqnarray} where $k_{\pm} = \sfrac{1}{2} q \pm k$, and the $\Lambda^{ijk}_{\mu}$ are as defined in \Eq{VPPvertex}. Isospin invariance entails that the photon couples only to the charged $\pi$ meson in the above $\pi$-loop integration. The $\pi$-meson EM transition amplitude, given in terms of the usual pion EM form factor, is \begin{equation} \Gamma_{\mu}(k_+\+\sfrac{1}{2}Q;Q) = 2 \left(k_+\+\sfrac{1}{2}Q\right)_{\mu} \; F_{\pi}(Q^2) , \label{GammaPiPi} \end{equation} where $k_+\+\sfrac{1}{2}Q$ is the average momentum of the incoming and outgoing pions, and $Q=-q-q'$ is the momentum transfered to the final vector meson by the photon. Substituting the relation in \Eq{GammaPiPi} into \Eq{Lam2}, one obtains \begin{eqnarray} \Lambda_{\mu\alpha\beta}(q,q') &=& 8ig_{\rho\pi\pi}^2\!\! \int \!\! \frac{d^4\!k}{(2\pi)^4} \Delta(k_{+}) (k_{+}\+\sfrac{1}{2}Q)_{\mu} F_{\pi}(Q^2) \nonumber \\ &\times& \Delta(k_{+}\+Q) k_{\alpha} f^{VPP}(k_{-},k_{+}) \Delta(k_{-}) \nonumber \\ &\times& (k\+\sfrac{1}{2}Q)_{\beta} f^{VPP}(k_{+}\+Q,k_{-}) . \label{ArhorhoAmp} \end{eqnarray} As in the case of the vector meson self-energy $\Pi_{\mu\nu}(q)$, the $\pi$-meson loop integration in \Eq{ArhorhoAmp} necessarily samples momenta for which {\em both} pions are on-mass-shell when $q^2 = q^{\prime 2} = - m_{\rho}^2$. It follows that $\rho$-meson EM form factors are, in general, complex functions of $Q^2$. The imaginary parts of these form factors are associated with interference between the {\em photo-stimulated} decay of the $\rho$ meson into two pions and the usual decay of the $\rho$ meson into two pions. The real parts of the EM form factors correspond to the situation where the intermediate mesons remain off-mass-shell. Hence, they are not observed as an asymptotic state and the $\rho$ meson remains intact. The real parts of the EM form factors are associated with the charge distribution of the $\rho$ meson. Thus, only the real parts of the $\rho$-meson electromagnetic form factors are of interest to the present study. As in the previous calculation of the vector-meson self energy $\Pi_{\mu\nu}(q^2)$, a Feynman parametrization is introduced, thereby combining the three denominators of the pion propagators in \Eq{ArhorhoAmp} into a single denominator, \begin{eqnarray} \lefteqn{ \Delta(k_{+}) \Delta(k_{+}\-Q) \Delta(k_{-}) = \sfrac{1}{2} \!\! \int_{-1}^{+1} \!\!\!\! d\alpha_1 d\alpha_2 \bigg[k^2 + m_{\pi}^2 + \frac{q^2}{4} } \nonumber \\ &-& \alpha_1 k\!\cdot\!q - (\alpha_1+\alpha_2)(k\! \cdot\! Q + \sfrac{1}{2} q\! \cdot\! Q + \sfrac{1}{2} Q^2) \bigg]^{-3} . \end{eqnarray} The EM form factors of the vector meson are Lorentz invariant and can be calculated in any frame. Herein, the calculation is performed in the Breit frame for which $Q_{\mu} = (0,0,Q,0)$ and $q_{\mu}=(0,0,-\sfrac{1}{2} Q, i \sqrt{\sfrac{1}{4}Q^2 + m_{\rho}^2})$, $k\cdot Q = x \sqrt{1 - z^2} \sqrt{k^2 Q^2}$, $k\cdot q = -\frac{1}{2} k\cdot Q $ + $i z \sqrt{k^2(\sfrac{1}{4}Q^2 + m_{\rho}^2)}$, and the elastic condition $2 Q\cdot q = - Q^2$ is satisfied. Then using the projection operators $O^{[i]}_{\mu\alpha\beta}$ discussed above and translating the $k^2$-integration domain by \begin{equation} k \rightarrow \tilde k = k + \sfrac{1}{2} \alpha_1 q + \sfrac{1}{2}(\alpha_1 + \alpha_2) Q , \end{equation} one finds that the two-pion contribution to the $\rho$-meson EM form factor is given by \begin{eqnarray} G^{\pi\pi}_i(Q^2) &=& \frac{2i}{(2\pi)^3} \int_0^{\infty} \!\!\! k^2 dk^2 \!\!\! \int_{-1}^{+1} \!\!\!\!\! \sqrt{1\-z^2} dz dx d\alpha_1 d\alpha_2 \nonumber \\ & &\times \frac{ {\cal F}^i}{ [k^2 - k^2_0 - i \epsilon ]^3 }, \label{Gpipik} \end{eqnarray} where \begin{eqnarray} k_0^2 &=& - \sfrac{1}{4}(1\-\alpha_1^2)q^2 \- m_{\pi}^2 \+ \sfrac{1}{4}(\alpha_1\+\alpha_2)(\alpha_2\+1)Q^2 , \\ {\cal F}^{i} &=& O^{[i]}_{\mu \alpha \beta} \big[k+\sfrac{1}{2}(\alpha_1+1)q + \sfrac{1}{2}(\alpha_1+\alpha_2+1)Q \big]_{\mu} k'_{\alpha} k^{''}_{\beta} \nonumber \\ & & \times F_{\pi}(Q^2) g_{\rho\pi\pi}^2 f^{VPP}(\tilde{k}_{-},\tilde{k}_{+}) f^{VPP}(\tilde{k}_{+}\+Q,\tilde{k}_{-}) \\ k^{\prime} &=& k \+ \sfrac{1}{2} \alpha_1 q \+ \sfrac{1}{2}(\alpha_1\+\alpha_2)Q, \\ k^{\prime\prime} &=& k^{\prime} + \sfrac{1}{2} Q \\ \tilde{k}_{\pm} &=& \pm \left[ k + \sfrac{1}{2} \alpha_1 q + \sfrac{1}{2}(\alpha_1 + \alpha_2) Q \right]+ \sfrac{1}{2} q . \end{eqnarray} Performing the $k$ integration by parts twice, \Eq{Gpipik} yields \begin{eqnarray} \lefteqn{ G^{\pi\pi}_i(Q^2) = \frac{i}{(2\pi)^3} \int_{-1}^{+1} \!\!\!\! \sqrt{1\- z^2} dz \, dx \, d\alpha_1 \, d\alpha_2 } \nonumber\ \\ & & \times \bigg( \int_0^{\infty} \!\!\!\! dk^2 \frac{1}{ [k^2 \- k^2_0 \- i \epsilon ] } \big[\frac{\partial\;}{\partial k^2}\big]^2 \; k^2 {\cal F}^i - \frac{ {\cal F}^i }{ k^2_0 }\big|_{k^2 = 0} \bigg). \label{Gpipik2} \end{eqnarray} The {\em real} part of this amplitude is of interest, which is obtained from the principal parts of the integrations. As in the calculation of the vector-meson self energy $\Pi_{\mu\nu}(q)$ described in previous sections, calculation of the real part of such amplitude requires a knowledge of the off-mass-shell transition form factor $f^{VPP}(p_1,p_2)$ appearing in ${\cal F}^i$. This is the same form factor that appears in \Eq{FPP} for the calculation of the vector-meson self energy $\Pi^{PP}(q^2)$. The determination of this form factor is carried out in \Sec{Sec:Model}, and the results of numerical calculations of the $\rho$-meson EM form factors will be postponed until \Sec{Sec:Results}. However, it is important to note that the fact that $f^{VPP}(p_1,p_2)$ appears in the calculations of both $\Pi_{\mu \nu}(q)$ and $G_E(Q^2)$ provides a self-consistency check of the dynamical model for $f^{VPP}(p_1,p_2)$. For a dynamical model to be acceptable, it must lead to an off-mass-shell transition form factor $f^{VPP}(p_1,p_2)$ that when substituted into \Eqs{PiPP} and (\ref{Gpipik}) gives reasonable results for both the $\rho$-$\omega$ mass splitting and the $\rho$-meson electromagnetic form factor. This issue is discussed further in \Sec{Sec:Results}. Since the $\rho$-meson EM form factor receives contributions from its quark substructure as well as from meson loops, the $\rho$ meson electric-monopole form factor is written as the sum: \begin{equation} G_{E}(Q^2) = \frac{G^{q\bar{q}}_{E}(Q^2) + G^{\pi\pi}_{E}(Q^2)} {G^{q\bar{q}}_{E}(0) + G^{\pi\pi}_{E}(0)} . \label{Ge} \end{equation} Here, the factors in the denominator ensure the unit charge of the $\rho$ meson $G_E(0) = 1$. In \Ref{Hawes}, it was noted that the quark-photon vertex obeys the Ward identity which results from the $U(1)$-gauge invariance of QED. When quark-photon vertex satisfies this identity and the Bethe-Salpeter amplitudes for the $\rho$ meson are normalized in the canonical manner, the charge conservation condition [$G^{\bar{q}q}_E(0) = 1$] for the $\rho$ meson is guaranteed. A similar Ward identity constrains the behavior of the pion-photon vertex, \begin{equation} \lim_{Q\rightarrow 0} \Gamma_{\mu}(p,p-Q) = \frac{d\;}{d p_{\mu} } \, \Delta^{-1}(p) . \label{WardId} \end{equation} This identity is satisfied by the form introduced in \Eq{GammaPiPi}. It is straight-forward to show that this identity provides a relation between the magnitude of the form factor $G_E^{\pi\pi}(0)$ and the derivative of the self energy $\Pi^{PP}(-m_{\rho}^2,m_{\pi},m_{\pi})$. This is done by examining the amplitude $\Lambda_{\mu\alpha\beta}(q,q')$ of \Eq{ArhorhoAmp} in the limit that the photon momentum $Q \rightarrow 0$. Then, with the identity in \Eq{WardId}, one can show that \begin{eqnarray} \lefteqn{ \lim_{Q\rightarrow 0} O^{[E]}_{\mu\alpha\beta} \Lambda_{\mu\alpha\beta}(q,Q\-q) } \nonumber \\ &=& O^{[E]}_{\mu\alpha\beta} \, 2 q_{\mu} {\bf T}_{\alpha\beta}(q,\-m_{\rho}^2) \left( \frac{d\;}{dq^2} \Pi^{PP}(q^2,m_{\pi},m_{\pi}) \right)_{\!\!q^2=-m_{\rho}^2} \!\!\!, \!\! \nonumber \\ &=& \delta Z_{\pi\pi}, \nonumber \end{eqnarray} where $\delta Z_{\pi\pi}$ is that part of the contribution to the normalization of the $\rho$ meson that arises from the two-pion intermediate state: \begin{equation} \delta Z_{\pi\pi} = \bigg( \frac{d\;}{dq^2} \Pi^{PP}(q^2,m_{\pi},m_{\pi}) \bigg)_{q^2 = -m_{\rho}^2} \label{deltaZpipi} . \end{equation} It follows immediately that $G_{E}^{\pi\pi}(0) = \delta Z_{\pi\pi}$, which is the justification of the choice of normalizations in \Eq{Ge}. From \Eq{Ge}, one may calculate the $\rho$ meson EM charge radius in the usual manner, \begin{eqnarray} \langle r^2 \rangle &=& - 6 \; \frac{d}{dQ^2} G_{E}(Q^2) \\ &=& \frac{-6}{1 + G_{E}^{\pi\pi}(0)} \bigg( \frac{d}{dQ^2} G_{E}^{\bar{q}q}(Q^2) \+ \frac{d}{dQ^2} G_{E}^{\pi\pi}(Q^2) \bigg) .\!\! \label{RadiusDef} \end{eqnarray} The charge radius that results from a numerical calculation of \Eq{RadiusDef} using \Eq{Gpipik2} is given in \Sec{Sec:Results}. \section{Approximate Reduction to Time-Ordered Contributions} \label{Sec:Orders} The approach to evaluating the meson-loop corrections to the $\rho$-$\omega$ mass splitting used here is to evaluate the covariant Feynman diagram of Fig.~\ref{Fig:Loop}, which includes all possible time orderings of the vertices and intermediate meson propagators. This is in contrast with the work of Ref.~\cite{GI}, which evaluates the mass splitting using time-ordered perturbation theory (TOPT), keeping only that part of the Feynman diagram corresponding to two mesons propagating forward in time (the forward-forward time ordering). In TOPT there is a separate contribution from the two-meson Z graph, where the vertices are ordered in such a way that the initial process (for, say, the $\rho$ self energy due to two-pion loops) is the creation of $\rho\pi\pi$ from the vacuum, which then propagate forward in time with the original $\rho$ meson until the original $\rho$ meson annihilates with the two $\pi$ mesons. This is the backward-backward time ordering. A calculation of the forward time ordering requires knowledge of the $\rho\pi\pi$ vertex at and near the vicinity of the real decay kinematics, where it is reasonably well known. However, a calculation of the backward time ordering requires knowledge of the vertex for the vacuum to $\rho\pi\pi$ process, which is not well known. Fortunately, the vertex for $\rho\pi\pi$ creation and annihilation is believed to be strongly suppressed, and the backward-backward diagram is further suppressed by the energy denominators of the $\rho\rho\pi\pi$ intermediate state propagation. For these reasons, the Z-diagram contribution is excluded from the calculation described in Ref.~\cite{GI}. With the belief that the multiple-hadron creation and annihilation vertices should suppress the backward-backward time-ordered diagrams, a simple calculation is undertaken to assess the relative importance of this diagram to the forward-forward time ordering in the present covariant framework. One might argue that predictions obtained within this covariant framework would be neither robust nor reliable if the forward-forward and backward-backward contributions were both large, but of opposite sign and canceled each other. Such subtle cancellations would be very sensitive to details of the model which may not be constrained accurately enough to be trusted. If there were a cancellation between the forward-forward and backward-backward diagrams in the present framework, it could explain the discrepancy between the small vector-meson mass shift obtained herein, and the larger mass shifts obtained in the study of \Ref{GI}, which keeps only the forward-forward contributions. To determine the relative importance of the various time-ordered contributions, one starts with the expression in \Eq{PiPP}, and then rewrites the $\rho$-meson self energy as a sum of four contributions arising from the four possible time orderings of the two pion propagators. One of these terms is identified with the contribution to the self energy due to the propagation of two {\em forward-propagating} (positive-energy) pions. The analytic structure and physical interpretation of this term, which is analogous to that calculated in \Ref{GI}, and the three other terms is discussed in detail. In \Sec{Sec:ResultsOrders}, a direct numerical evaluation of these terms will show that for time-like values of the vector-meson four momentum $q$, the dominant contribution to the self energy $\Pi^{PP}(q^2)$ comes from the term identified with the propagation of two forward-propagating pions. Therefore, although the inclusion of {\em all} four time orderings is necessary to maintain the Lorentz covariance of the present framework, the additional time-orderings present in covariant expressions like \Eq{PiPP}, are {\em not} responsible for the significant discrepancy between the mass shifts obtained herein (to be presented in \Sec{Sec:ResultsOrders}) and those obtained by \Ref{GI}. Rather, in the present quantum field theoretic framework, one observes that the term in vector-meson self energy that arises from two forward-propagating pions is, in fact, the dominant contribution. Before proceeding further, it is necessary to clarify some issues concerning the use of quantum field theoretic models formulated in Euclidean space. In the present study, observables are obtained by analytically continuing the arguments of the Schwinger functions (Euclidean Green functions) into Minkowski space. In principal, this is done by first integrating over the {\em internal} momenta so that the Schwinger function depends only on Lorentz-scalar products of the external momenta. The Schwinger function can then be analytically continued in these scalar products into Minkowski space. For example, once the integration over the internal momenta $k$ is carried out in \Eq{PiPP}, the resulting Schwinger function $\Pi^{PP}(q^2)$ depends only on the square of the vector-meson four momentum $q^2$. The Schwinger function is then analytically continued into Minkowski space by letting $q^2 \rightarrow - m_{\rho}^2$.\footnote{ Additional details concerning the formulation of quantum field theory in Euclidean space and the analytic properties of Schwinger functions may be found in \Ref{Wightman}.} In the present study, the use of model forms for the elementary Schwinger functions (e.g., quark propagators, Bethe-Salpeter amplitudes and meson transition vertices) allows the analytic continuation of external momenta to be performed in a straight-forward manner by letting {external} momenta become complex before the {internal} momenta are integrated over. An important ramification of such a framework is that a direct comparison between the elementary Schwinger functions employed herein and those employed in other models formulated in Minkowski space is impossible. One reason for this is that the model form of the quark propagators $S_f(k)$, given by \Eqs{QuarkProp}, are parametrized in terms of {\em entire} functions of $k$. This parametrization provides the model with quark confinement by ensuring that the quark propagator $S_f(k)$ is free of singularities in the finite, complex-$k$ plane. Hence, the quark propagator has no Lehman representation and no quark-production thresholds can arise from the calculation of observables. Of course, the use of such quark propagators also forbids the possibility of performing a Wick rotation in the loop-integration variable $k$. Therefore, the elementary Schwinger functions employed herein are permanently embedded in Euclidean space. Only final quantities, such as $\Pi^{PP}(q^2)$, can be analytically continued into Minkowski space. Another feature of the framework that makes a Wick rotation impossible to perform is the rich analytic structure of the meson transition vertices, e.g., $\Lambda^{ijk}_{\mu}(p,p')$. A Wick rotation of these transition vertices, although possible in principal, is made difficult by the presence of singularities which arise from the many-particle nature of the field theory. In general, these singularities may cross the integration path requiring that exceptional care be taken when performing a Wick rotation. In practice, these transition vertices are calculated and parametrized only on the small domain of the complex-momentum plane for which they are needed in pion-loop integrations; their behavior outside this domain is unknown. Such difficulties are not present in a quantum mechanical formulation that truncates the Fock space in such a way as to include only states with a small number of particles. The analytic properties of transition vertices obtained in such models are more easily analyzed. Nevertheless, it is possible to make a simple connection between this field theoretic approach and TOPT. The starting point is the calculation of the $\rho$-meson self energy due to the two-pion loop from \Eq{PiPP}. As in \Ref{GI}, the calculation is carried out in the rest frame of the $\rho$ meson. In this frame, $q_{\mu} = (\vec{0},im_{\rho})$ and the $\pi$-meson propagators, which depend on the momenta $k_{\pm} = k \pm \sfrac{1}{2} q$, can be written in the following form \begin{eqnarray} \Delta(k_{\pm},m_{\pi}) &=& \frac{-1}{ (ik_4 \mp \frac{1}{2}m_{\rho})^2 - (\omega(\vec{k}^2) - i \epsilon)^2} \label{PiProp1}, \end{eqnarray} where $\omega(\vec{k}^2) = \sqrt{\vec{k}^2 + m_{\pi}^2}$. The denominator of \Eq{PiProp1} is the difference of two squares. It can be factorized and rewritten as a difference of two poles in the complex-$k_4$ plane, \begin{eqnarray} \Delta(k^2_{\pm},m_{\pi}) = \frac{1}{2 \omega(\vec{k}^2)} &\bigg[& \frac{1}{ik_4 \mp \frac{1}{2}m_{\rho} + \omega(\vec{k}^2) - i \epsilon} \nonumber \\ & & - \frac{1}{ik_4 \mp \frac{1}{2}m_{\rho} - \omega(\vec{k}^2) + i \epsilon} \bigg] \label{TOProp}. \end{eqnarray} In a field theory based in Minkowski space, the two terms in \Eq{TOProp} would be identified with forward- and backward-propagating pions. A similar identification is possible for the two-pion system in our approach by substituting \Eq{TOProp} into \Eq{PiPP}. One then obtains an expression for the vector meson self energy in which four poles in the complex-$k_4$ plane are explicit. Writing the self energy as a sum of four terms: $\Pi^{PP}(q^2)=\Pi^{[++]}(q^2)+\Pi^{[--]}(q^2) +\Pi^{[+-]}(q^2)+\Pi^{[-+]}(q^2)$, where, for $a, b = +,-$, one has \begin{eqnarray} \Pi^{[ab]}(-m_{\rho}^2) = -\frac{4}{3} \int\! \frac{d^4k}{(2\pi)^4} \frac{{\cal F}^{PP}(k,q) }{4\omega^2(\vec{k}^2)} \; A^{[ab]}(k) \label{TOPiPP}, \end{eqnarray} and \begin{eqnarray} A^{[++]}(k) &=& \left(\frac{1}{ik_4-\frac{1}{2}m_{\rho}+\omega(\vec{k}^2)-i\epsilon}\right) \nonumber \\ &&\times \left(\frac{-1}{ik_4+\frac{1}{2}m_{\rho}-\omega(\vec{k}^2)+i\epsilon} \right) , \label{A++} \\ A^{[--]}(k) &=& \left(\frac{-1}{ik_4-\frac{1}{2}m_{\rho}-\omega(\vec{k}^2)+i\epsilon}\right) \nonumber \\ && \times \left(\frac{1}{ik_4+\frac{1}{2}m_{\rho}+\omega(\vec{k}^2)-i\epsilon}\right) , \\ A^{[+-]}(k) &=& \left(\frac{1}{ik_4-\frac{1}{2}m_{\rho}+\omega(\vec{k}^2)-i\epsilon}\right) \nonumber \\ && \times \left(\frac{1}{ik_4+\frac{1}{2}m_{\rho}+\omega(\vec{k}^2)-i\epsilon}\right) , \\ A^{[-+]}(k) &=& \left(\frac{1}{ik_4-\frac{1}{2}m_{\rho}-\omega(\vec{k}^2)+i\epsilon}\right) \nonumber \\ && \times \left(\frac{1}{ik_4+\frac{1}{2}m_{\rho}-\omega(\vec{k}^2)+i\epsilon} \right) . \end{eqnarray} These are identified as contributions to the $\rho$-meson self energy due to forward-forward ($++$), backward-backward ($--$), forward-backward ($+ -$) and backward-forward ($-+$) propagation of the two pions, respectively. This nomenclature specifies the four possible time orderings of the two pions in the pion loops. It is justified from an analysis of the motion of the poles in the complex-$k_4$ plane as the three-momentum $\vec{k}$ is varied. \begin{figure}[ht] \begin{center} \epsfig{figure=./MAPnew.eps,width=6.0cm,angle=-90} \end{center} \caption{ The positions in the complex-$k_4$ plane of the poles in \Eq{TOPiPP}. Dashed lines with arrows indicate the motion of these poles as the relative three-momentum $|\vec{k}|$ is {\em decreased} from infinity to zero. The positions of the poles for $|\vec{k}| = 0$ are denoted by ``X.'' The diagram on the left depicts the fictional situation for which $m_{\rho} \approx m_{\pi}/2 < 2 m_{\pi}$ and the $\rho \rightarrow \pi\pi$ decay is forbidden by energy conservation. The diagram on the right depicts the situation for which $m_{\rho} > 2 m_{\pi}$ and the $\rho \rightarrow \pi \pi$ decay is allowed. For a certain values of $|\vec{k}|$, two of the pion poles may pinch the $k_4$-contour integration as $\epsilon \rightarrow 0$. This signals the threshold of the $\rho \rightarrow \pi\pi$ decay and results in a cut in the self energy of the $\rho$ meson. These two poles correspond to the forward-forward time ordering in \Eq{A++}. } \label{Fig:Poles} \end{figure} These integrands in \Eq{TOPiPP} exhibit poles for particular values of the fourth component of the loop-integration momentum $k_4$. The positions of these poles depend on the magnitude of the three-momentum $\vec{k}$ as well as the mass of the vector meson $m_{\rho}$. In \Fig{Fig:Poles}, the positions of these poles, along with their motions as $|\vec{k}|$ is varied, are depicted for two possible scenarios. The left diagram corresponds to a scenario (not realized in nature) for which the $\rho$-meson mass $m_{\rho} \approx m_{\pi}/2 < 2 m_{\pi}$. In this situation, the center-of-momentum energy of the $\rho$ meson is below the two-pion decay threshold; hence, the $\rho$ meson is stable with respect to the strong decay into two pions. The diagram on the right depicts the situation (realized in nature) for which the mass of the $\rho$ meson $m_{\rho} > 2 m_{\pi}$ and the strong, $\rho \rightarrow \pi \pi$ decay is allowed. In both diagrams, the motions of these poles as $|\vec{k}|$ is varied are denoted by dashed lines. Arrows depict the trajectory of the poles as $|\vec{k}|$ is {\em decreased} from infinity to zero. The positions of the poles for $|\vec{k}| = 0$ are denoted by ``X.'' Examination of the left diagram in \Fig{Fig:Poles} depicts the fictional situation where the $\rho$-meson mass is below the $\rho\rightarrow\pi\pi$ threshold energy. Hence, the integrations over $k_4$ in \Eq{TOPiPP} never encounter poles for any value of $|\vec{k}|$. Thus, in this scenario, the $\rho$-meson self energy $\Pi^{PP}(q^2)$ is purely real and the $\rho$ meson is stable with respect to strong interactions. In the right diagram of \Fig{Fig:Poles} for which $m_{\rho} > 2 m_{\pi}$, one observes a very different situation. Here, as the relative momentum $|\vec{k}|$ is decreased from $-\infty$, two of the poles pass each other, thereby causing the contour integration path for $k_4$ to become ``pinched'' in the limit that $\epsilon \rightarrow 0$. Such a pinching of the $k_4$-integration contour results in a branch cut in $\Pi^{PP}(q^2)$ and hence a non-zero imaginary part of the vector-meson self energy. This signals the opening of the two-pion decay channel and provides the $\rho$ meson with a finite two-pion decay width. The value of $|\vec{k}|$ for which this pinch first develops is obtained from the energy-conservation condition: $\omega(\vec{k}^2)=m_{\rho}/2$. The fact that this pinch singularity in $\Pi^{PP}(q^2)$ can only arise in the calculation of $\Pi_{++}(q^2)$ is the justification for this term being identified with the propagation of two forward-propagating pions. Clearly, only positive-energy pions may appear in asymptotically free states and in so doing contribute to the imaginary part of the vector-meson self energy. The term $\Pi^{--}(q^2)$ is identified as arising from two-backward propagating pions by noting that had one started with the convention that energies were negative rather than positive quantities, the replacement of the $\rho$-meson center-of-momentum energy $m_{\rho}$ with $-m_{\rho}$, would result in the pinch singularity appearing in $\Pi^{--}(q^2)$ rather than in $\Pi^{++}(q^2)$. That is, within a framework that employed such a convention, the asymptotically free states would necessarily contain negative-energy pions. A corollary that follows from the above argument is that in the limit that $q^2 \rightarrow 0$, no energy flows into the two-pion-loop diagram and there is no difference between the forward-forward and backward-backward time orderings; that is, for a massless vector meson these time orderings contribute equally to its self energy. (This is indeed what is observed numerically, as discussed in \Sec{Sec:ResultsOrders}.) It is important to note that the contribution to the vector-meson self energy due to negative energy pions does indeed appear in quantum mechanics, even though there are no negative energy states in the Fock space. By considering the time-ordered Feynman diagram associated with $\Pi^{--}(q^2)$, written so that all particles move in the positive-time direction, one observes that the four-particle Fock state $|\rho\rho\pi\pi\rangle$ has a non-zero overlap with the intermediate particles depicted in the diagram. This state contributes only to the real part of the vector-meson self energy because energy conservation forbids the decay $\rho \rightarrow \rho\rho\pi\pi$. In a quantum mechanical framework, such as that employed in \Ref{GI}, Cauchy's Theorem entails that the mixed, forward-backward ($+-$) and backward-forward ($-+$) contributions, are identically zero. However, in the present framework the meson-transition vertex in \Eq{fVPPForm} and quark propagator in \Eq{QuarkProp} have essential singularities at infinity. Therefore, one cannot close the contour the integration at infinity and use Cauchy's theorem to show that these terms are zero. Consequently, there is no such constraint on the value of these terms in the present approach, and they will generally be different from zero. A discussion of the physical interpretation of these terms and the role they play in determining the unitarity of the theory is left for future investigations. It is sufficient for the present application to note that in \Sec{Sec:ResultsOrders}, a direct numerical evaluation of these terms shows them to provide a negligible contribution to the vector-meson self energy. Hence, they have no impact on the results presented herein. \section{Meson transition amplitudes} \label{Sec:Model} In the previous sections, expressions for the vector-meson self energy and EM form factors were given in terms of meson propagators and meson transition amplitudes. In the following section, these transition amplitudes are given in terms of nonperturbative quark-loop integrations that depend on model Bethe-Salpeter amplitudes and nonperturbatively-dressed quark propagators developed in studies of hadron observables based on the Schwinger-Dyson equations of QCD. A recent review of such studies is provided by \Ref{Roberts}. The imaginary part of the vector-meson self energy is determined by meson transition amplitudes for which all external meson momenta are on-mass-shell. It follows that an experimental determination of decay widths for the vector meson provides only the {\em on-mass-shell} values of the meson transition amplitudes. In the following, the values of the transition amplitudes when all mesons are on-mass-shell are referred to as {\em coupling constants}, e.g., $g_{\rho\pi\pi}$. A determination of the real parts of the vector meson self energy requires calculating the principal part of the integrals in \Eqs{PiPP} and (\ref{PiVP}), which sample the $k^2$-integration domain for which the intermediate mesons are off-mass-shell. Thus, a calculation of the real parts of the self energy and EM form factors requires knowledge of the meson transition form factors for off-mass-shell values of the intermediate meson momenta. As the real part of the self energy is not directly observable, experiment cannot determine the behavior of the off-mass-shell-meson transition form factors. Nonetheless, experimental observations can provide some guidance in helping to constrain the real parts of self energies and EM form factors. This can be done by looking for observables which are proportional to the {\em difference} of two self energies. For example, from \Eqs{PiSum:Rho} and (\ref{PiSum:Omega}), it is apparent that the $\rho$--$\omega$ mass splitting provides a measure of the difference of the real parts of the $\rho$ and $\omega$ self energies. In \Sec{Sec:Model}, the calculation of off-mass-shell meson transition form factors is discussed. The VVP and VPP vector-meson transition amplitudes are calculated within a generalized impulse approximation in which the mesons couple to each other by means of a nonperturbatively-dressed quark loop. In this generalized impulse approximation, the quark propagators and meson Bethe-Salpeter amplitudes are nonperturbatively dressed, but explicit three-body interactions are neglected. In the following, quark-loop integrations are performed in Euclidean space (where $q^2 > 0$), and final results are analytically continued to Minkowski space by letting the external (timelike) momenta $q^2 \rightarrow - m^2_{V}$. The Dirac matrices $\gamma_{\mu}$ employed herein are Hermitian, $\gamma_{\mu}^{\dag} = \gamma_{\mu}$, and satisfy the anti-commutation relation $\{ \gamma_{\mu},\gamma_{\nu} \} = 2 \delta_{\mu \nu}$. The amplitude describing the coupling of a vector meson of momentum $q = -p_1-p_2$ and two pseudoscalar mesons with momentum $p_1$ and $p_2$ is given in impulse approximation by\footnote {The sense of all momenta is such that positive momenta flow {\em into} vertices.}, \begin{eqnarray} \lefteqn{ \Lambda^{ijk}_{\mu}(p_1,p_2) = {\rm tr}_{CFD}\int \! \frac{{\rm d}^4k}{(2\pi)^4} {\bf S}(k_{++}) \lambda_{i} V_{\mu}(k\+\sfrac{1}{2}p_2;q) } \nonumber \\ & & \times {\bf S}(k_{-+}) \lambda_{j} \Gamma_{P}(k\-\sfrac{1}{2}q;p_1) {\bf S}(k_{--}) \lambda_{k} \Gamma_{P}(k;p_2) , \label{LamVPP:Quark} \end{eqnarray} where $k_{\alpha\beta} = k + \sfrac{\alpha}{2} q + \sfrac{\beta}{2} p_2$, $\lambda_i$ are linear combinations of the Gell-Mann SU(3)-flavor matrices associated with the mesons involved, the trace is over color, flavor and Dirac indices, and ${\bf S}(k) = {\rm diag}(S_u(k),S_d(k),S_s(k))$ is the quark propagator for $u$, $d$ and $s$ quarks. It is a 3$\times$3 matrix in the quark-flavor space. In this study, the effects of SU(3)-flavor breaking are neglected, so the quark propagator is diagonal in flavor indices and $S_u(k) = S_d(k) = S_s(k)$. Here $V_{\mu}(k,k')$ is the vector-meson Bethe-Salpeter amplitude for the $\rho$ or $\omega$ mesons, and $\Gamma_{P}(k;p_1)$ is the Bethe-Salpeter amplitude for the $P$ = $\pi$ or $K$ mesons. The flavor structure of the meson Bethe-Salpeter amplitudes is given explicitly in \Eq{LamVPP:Quark} in terms of the 3$\times$3 Gell-Mann matrices $\lambda_i$ in the quark-flavor space. The meson transition amplitude $\Lambda_{\mu}^{ijk}(p_1,p_2)$ is depicted in \Fig{Fig:Loop} as a triangle diagram. The transition form factor defined by \Eq{VPPvertex} is obtained from \Eq{LamVPP:Quark} by choosing values of the flavor indices $i$, $j$ and $k$ appropriate for the particular process $V\rightarrow PP$ under consideration, and then contracting the amplitude with an appropriate vector. In the case of $\rho^{0} \rightarrow \pi^{+} \pi^{-}$, one obtains \begin{equation} g_{\rho\pi\pi} f^{VPP}(p_1,p_2) = \frac{ (p_1 + p_2)_{\mu}}{m_V^2 - 4 m_P^2} \Lambda^{3-+}_{\mu}(p_1,p_2) \label{fVPP_Proj} , \end{equation} where the flavor indices refer to the following linear combinations of Gell-Mann matrices $\lambda_{\pm} = \sfrac{1}{\sqrt{2}}( \lambda_1 \pm i \lambda_2)$, which are associated with an isovector meson of good charge. The resulting form factor $f^{VPP}(p_1,p_2)$ is the essential dynamical element necessary to calculate the real part of the two-pseudoscalar intermediate state contribution to the vector-meson self energy $\Pi^{PP}(q^2)$ from \Eq{PiPP}. Transition amplitudes for processes involving strange mesons, such as $\rho \rightarrow K \bar{K}$, are obtained by assuming SU(3)-flavor symmetry is exact, so that $S_s(k) = S_u(k)$ and $\Gamma_{K}(k;p_1) = \Gamma_{\pi}(k;p_1)$. From considerations of SU(3)-flavor symmetry, one can show that the left-hand-side of \Eq{fVPP_Proj} gains an additional factor of $1/\sqrt{2}$ for the process $\rho \rightarrow K \bar{K}$. In particular, for $\rho^+ \rightarrow K^+ \bar{K}^0$, the analogous result for \Eq{fVPP_Proj} would be \begin{equation} \frac{ g_{\rho\pi\pi}}{\sqrt{2}}f^{VVP}(p_1,p_2) = \frac{ (p_1 + p_2)_{\mu}}{m_V^2 - 4 m_P^2} \Lambda^{3 K^{-} {K}^0}_{\mu}(p_1,p_2) \label{fVPP_Proj2} . \end{equation} Here, the flavor indices denote the linear combinations of Gell-Mann matrices, $\lambda_{K^{-}} \equiv \sfrac{1}{\sqrt{2}} ( \lambda_4 - i \lambda_5)$ and $\lambda_{{K}^{0}} \equiv \sfrac{1}{\sqrt{2}} ( \lambda_6 + i \lambda_7)$. The additional factor of $1/\sqrt{2}$ that appears in \Eq{fVPP_Proj2} relative to that in \Eq{fVPP_Proj}, is the reason for the factor of $1/2$ that appears as the coefficient of $\Pi^{PP}(q^2,m_K,m_K)$ in \Eqs{PiSum:Rho} and (\ref{PiSum:Omega}). Similarly, the coupling of two vector mesons of momenta $-(p_1+p_2)$ and $p_1$ to a pseudoscalar meson of momentum $p_2$ is given in the impulse approximation by \begin{eqnarray} \lefteqn{ \Lambda^{ijk}_{\mu\nu}(p_1,p_2) = {\rm tr}_{CFD}\int \! \frac{{\rm d}^4k}{(2\pi)^4} {\bf S}(k_{++}) \lambda_{i} V_{\mu}(k\+\sfrac{1}{2}p_2;q) } \nonumber \\ & & \times {\bf S}(k_{-+}) \lambda_{j} V_{\nu}(k\-\sfrac{1}{2}q;p_1) {\bf S}(k_{--}) \lambda_{k} \Gamma_{P}(k;p_2) . \label{LamVVP:Quark} \end{eqnarray} The elements appearing in \Eq{LamVVP:Quark} are the same as in \Eq{LamVPP:Quark}. No new elements are necessary to obtain the $VVP$ transition amplitude. The form factor $f^{VVP}(p_1,p_2)$ is obtained from \Eq{LamVVP:Quark} using \Eq{VVPvertex} and is the essential dynamical element necessary to calculate the real part of the intermediate vector-pseudoscalar contribution to the vector-meson self energy $\Pi^{VP}(q^2,m_V,m_P)$ from \Eq{PiVP}. Here too, the assumption of exact-SU(3)-flavor invariance provides a simple relation between the processes $\rho \rightarrow \omega \pi$, $\rho \rightarrow \rho \eta$, $\rho \rightarrow K^* K$, $\omega \rightarrow \rho \pi$, $\omega \rightarrow \omega \eta$, and $\omega \rightarrow K^* K$. The flavor factors that relate these processes lead to the coefficients appearing in front of the vector-pseudoscalar contributions in \Eqs{PiSum:Rho} and (\ref{PiSum:Omega}). Evaluation of the quark-loop integrations in \Eqs{LamVPP:Quark} and (\ref{LamVVP:Quark}) requires knowledge of the $u$-, $d$- and $s$-quark propagators as well as the vector-meson and pseudoscalar-meson Bethe-Salpeter amplitudes. These are taken from phenomenological studies of the electromagnetic form factors and strong and weak decays of pseudoscalar and vector mesons \cite{Hawes,Burden}. The form of the model, dressed-quark propagators are based on several numerical studies of the quark Schwinger-Dyson equation using a realistic model-gluon propagator \cite{Frank,Maris}. In a general covariant gauge, the propagator for a quark of flavor $f$ is written as $S_f(k) = - i \gamma \cdot k \sigma_V^f(k^2) + \sigma_S^f(k^2)$. It is well described by the following parametrization: \begin{eqnarray} \bar{\sigma}^f_S(x) & = & \xi[b^f_1 x] \; \xi[b^f_3 x] \left(b^f_{0} + b^f_2 \xi[ \Lambda x] \right) \nonumber\\ & & + 2 \bar{m}_f \xi\big[ 2(x+\bar{m}_f^2) \big] + C_{f} e^{-2x}\:, \nonumber\\ \bar{\sigma}^f_V(x) & = & \frac{ 2(x+\bar{m}_f^2) - 1 + e^{-2(x+\bar{m}_f^2)} } { 2(x+\bar{m}_f^2)^2 }\:, \label{QuarkProp} \end{eqnarray} where $\xi[x] = (1 - e^{-x})/x$, $x = k^2/\lambda^2$, $\bar{\sigma}^f_{S} = \lambda \sigma^f_S$, $\bar{\sigma}^f_V = \lambda^2 \sigma^f_V$, $\bar{m}_f = m_f / \lambda$, $\lambda =$ 0.566~GeV, $\Lambda = 10^{-4}$, and where the $b_i^f$, $C_{f}$ and $m_f$ are parameters. Since $\xi[x]$ is an entire function, this dressed-quark propagator has {\em no} Lehmann representation. The lack of a Lehmann representation for $S_f(k)$ is sufficient to ensure that the quarks are confined, since no quark-production thresholds can appear in the calculation of observables. The quark propagator $S_f(k)$ also reduces to a bare-fermion propagator when its momentum is large and spacelike, in accordance with the asymptotic behavior expected from perturbative QCD (up to logarithmic corrections). The parameters for the $u$-, $d$- and $s$-quark propagators were determined in \Ref{Burden} by performing a $\chi^2$ fit to a range of $\pi$- and $K$-meson observables. The parameters for the $s$-quark propagator employed herein were re-fit in \Ref{Hawes}, also from a $\chi^2$ fit to $K$-meson observables, but with the additional constraint that the resulting $s$-quark propagator more closely resemble that obtained by the numerical study of \Ref{Maris}. The resulting parameters are given in \Tab{Tab:QuarkParam}. \begin{table} \begin{center} \caption{Confined-quark propagator parameters for $u$, $d$, and $s$ quarks from \Ref{Hawes}. An entry of ``same'' for a meson indicates that the value of the parameter is the same as for the quark flavor with which it is grouped.} \begin{tabular}{l|lllllc} $f$/meson& $b^f_0$ & $b^f_1$ & $b^f_2 $ & $b^f_3$ & $C_{f}$ & $m_f$ [MeV]\\ \hline $u,d$ & 0.131 & 2.900 & 0.603 & 0.185 & $\;\;$0 & 5.1 \\ $\pi$ & same & same & same & same & 0.121 & $0\;\;\;$ \\ $\rho$, $\omega$& 0.044 & 0.580 & same & 0.462 & $\;\;$0 & $0\;\;\;$ \\ \hline $ s $ & 0.105 & 2.900 & 0.540 & 0.185 & $\;\;$0 & 130.0 \\ $ K $ & 0.322 & same & same & same & $\;\;$0 & $0\;\;\;$ \\ $K^*$ & 0.107 & 0.870 & same & 0.092 & $\;\;$0 & $0\;\;\;$ \\ \end{tabular} \label{Tab:QuarkParam} \end{center} \end{table} In the chiral limit $m_f \rightarrow 0$, Goldstone's theorem allows one to obtain the pseudoscalar-meson Bethe-Salpeter amplitude $\Gamma_{P}(k ; P)$ directly from the dressed-quark propagator $S_f(k)$ \cite{Maris}. For the calculation of infrared transition amplitudes, like those studied herein, it is sufficient to keep only the part of the pseudoscalar Bethe-Salpeter amplitude that is proportional to $\gamma_5$. In this case, the amplitude is given by \begin{equation} \Gamma_{P}(k ; P) = \gamma_5 \: \frac{ B_P(k^2) }{N_{P}} \:, \label{PiBSAmp} \end{equation} where $B_P(k^2)$ has the same form as $B_f(k^2)$, one of the Lorentz invariants appearing in the quark inverse propagator $S_f^{-1}(k) = i \gamma \cdot k A_f(k^2) + B_f(k^2)$, but with the parameters given in \Tab{Tab:QuarkParam} for $P = \pi$ or $K$. The $\pi$- and $K$-meson Bethe-Salpeter amplitudes are obtained from Eqs.~(\ref{QuarkProp}) and(\ref{PiBSAmp}) using the parameters from \Tab{Tab:QuarkParam}. In \Eq{PiBSAmp}, $N_P$ is the on-mass-shell normalization for the $P=$ $\pi$ or $K$ meson, and is determined by the normalization condition: \begin{eqnarray} 1 &=& \frac{ N_c p_{\mu}}{m_{P}} {\rm tr}_{D} \int\! \frac{{\rm d}^4 k}{(2\pi)^4} \nonumber \\ & & \frac{\rm d\;}{{\rm d}p_{\mu}} S(k\+\sfrac{1}{2}p) \Gamma_{P}(k;p) S(k\-\sfrac{1}{2}p) \bar{\Gamma}_{P}(k;p) \nonumber\ \\ & &+ S(k\+\sfrac{1}{2}p) {\Gamma}_{P}(k;p) \frac{\rm d\;}{{\rm d}p_{\mu}} S(k\-\sfrac{1}{2}p) \bar{\Gamma}_{P}(k;p) ,\label{fK} \end{eqnarray} where ${\rm tr}_{D}$ denotes a trace over Dirac indices and $N_c = 3$. In \Ref{Maris} it was shown that by maintaining only the Dirac amplitude proportional to $\gamma_5$ in the pion Bethe-Salpeter amplitude, one cannot reproduce the ultraviolet behavior of the pion electromagnetic form factor. To do so requires the inclusion of the axial-vector $\gamma_5 \gamma \cdot p$ components into the $\pi$-meson Bethe-Salpeter amplitude. Nonetheless, for the calculation of observables which involve small momentum transfers, neglecting Dirac moments other than $\gamma_5$ and normalizing the Bethe-Salpeter amplitude according to \Eq{fK} is usually sufficient. The vector meson Bethe-Salpeter amplitudes employed herein are taken from \Ref{Hawes}: \begin{equation} V_{\mu}(k; p) = \left( \gamma_{\mu} + \frac{p_{\mu} \gamma \cdot p}{m_V^2} \right) \frac{ B_V(k^2) }{N_{V}} \:, \label{VectorBSAmp} \end{equation} where $p$ is the momentum of the vector meson with mass $m_V$, $k$ is the relative momenta of the quark and antiquark, $N_V$ is the Bethe-Salpeter normalization, obtained from an obvious generalization of \Eq{fK}, and $B_V(k^2)$ has the same form as the Lorentz-invariant function $B_f(k^2)$ found from the quark propagator of \Eq{QuarkProp} with parameters given in \Tab{Tab:QuarkParam} with $V = \rho$, $\omega$ or $K^*$. Again, all but the leading Dirac moment proportional to $\gamma_{\mu}$ in the vector meson Bethe-Salpeter amplitude have been neglected. A numerical study \cite{BQRTT} of the Bethe-Salpeter equation that employed a separable model for the quark-antiquark scattering kernel found that the magnitude of this moment is about ten times larger than the next largest. However, some observables may be more sensitive to sub-leading Dirac moments in the Bethe-Salpeter amplitude. Indeed, this was the case observed in \Ref{Maris} for the asymptotic-$q^2$ behavior of the $\pi$ meson electromagnetic form factor. However, the simple form provided by \Eq{VectorBSAmp} is sufficient to provide a reasonable model for calculating the off-mass-shell transition form factors $f^{VPP}(p_1,p_2)$ and $f^{VVP}(p_1,p_2)$. With these elements having already been determined by previous studies, the on- and off-mass-shell values of the transition amplitudes $\Lambda^{VPP}_{\mu}(p_1,p_2)$ and $\Lambda^{VVP}_{\mu \nu}(p_1,p_2)$ can be calculated with no adjustable parameters. The value of a transition amplitude for on-mass-shell meson momenta is referred to as a ``coupling constant''. If the invariant mass of the final state $\sqrt{s}$ is less than the mass of the initial state, the decay is allowed to proceed and the coupling constant is experimentally observable. If energy conservation prevents all external mesons from being simultaneously on-mass-shell, one can still define the ``coupling constant'' as the value of the transition amplitude at this point. However, in this case, there is no decay from which it can be measured experimentally. In \Ref{Hawes}, the coupling constants $g_{\rho \pi \pi}$ and $g_{K^* K\pi}$ were obtained using the model quark propagators from \Eqs{QuarkProp} and Bethe-Salpeter amplitudes from \Eqs{PiBSAmp} and (\ref{VectorBSAmp}) with the parameters from \Tab{Tab:QuarkParam}. The values for $g_{\rho \pi \pi}$ and $g_{K^* K\pi}$ obtained therein were 8.52 and 9.66, respectively. The experimentally observed values for these coupling constants are $6.03\pm0.02$ and $6.40\pm0.04$, respectively. There are several possible reasons that this model gives results for the decay widths of a vector meson into two pseudoscalars that are larger than obtained by experiment. The first is that for this process in particular, higher Dirac moments contribute significantly to the magnitude of the coupling constant \cite{MAP2}. The second is the possibility that final-state interactions may also be important. For these reasons, rather than employ the value of $g_{\rho \pi \pi}$ obtained in \Ref{Hawes}, the experimentally determined value of $g_{\rho \pi \pi} = 6.03$ is used. The meson transition form factors $f^{VPP}(p_1,p_2)$ and $f^{VVP}(p_1,p_2)$ are the important dynamical element necessary for the present calculations. The scales of these are rather insensitive to details of the model Bethe-Salpeter amplitudes, so that the model of \Ref{Hawes} is quite adequate for their determination. The coupling constant $g_{\rho \omega \pi}$ = 10.81 is obtained from the $\rho\omega\pi$ transition amplitude in \Eq{LamVVP:Quark}, when all three mesons are on-mass-shell. Of course, the lack of phase space makes the decay $\omega \rightarrow \rho \pi$ impossible, so that the vector-meson self energy contribution due to vector-pseudoscalar intermediate meson states is real for a physical vector meson; i.e., the value of $g_{\rho \omega \pi}$ cannot be determined experimentally. Throughout this article, the value $g_{\rho \omega \pi}$ = 10.81 obtained from \Eqs{PiVP} with (\ref{FVP}) is used \cite{MAP3}. For comparison, the value of this coupling constant can been estimated from the anomalous Wess-Zumino term in an effective chiral Lagrangian with the assumption of a universal $\rho$-meson coupling. With the definition of the coupling constant from \Eq{VVPvertex}, it is found to be $g_{\rho \omega \pi} \approx 11.5$ \cite{UMeissner}, which is consistent with the results of the present calculation. Similarly, a study of the decays of the $\omega$-meson using vector-meson dominance finds the value $g_{\rho \omega \pi} \approx 9.7$~\Ref{Durso}. The loop integrations of Eqs.~(\ref{PiPP}) and~(\ref{PiVP}) sample the meson transition form factors $f^{VPP}(p_1,p_2)$ and $f^{VVP}(p_1,p_2)$ for {\em off-mass-shell} values of the intermediate-meson four momenta $p_1$ and $p_2$. A calculation of these form factors from the quark-loop integrations of \Eqs{LamVPP:Quark} and (\ref{LamVVP:Quark}) requires a suitable definition for meson Bethe-Salpeter amplitudes when their total momentum is off-mass-shell. Although Bethe-Salpeter amplitudes are uniquely determined from the homogeneous Bethe-Salpeter equation for on-mass-shell values of the total momentum, an extrapolation of the amplitude away from this on-mass-shell value is {\em not} uniquely determined, and hence, is model dependent. Herein, the Bethe-Salpeter amplitudes as defined in \Eqs{PiBSAmp} and (\ref{VectorBSAmp}) are used, even for values of the meson four momenta that are off-mass-shell. This is most similar to the procedure employed in quark model calculations such as \Ref{GI}, where a complete set of intermediate states is inserted between the vector-meson states leading to transition amplitudes for on-shell mesons. The off-shell behavior of the intermediate mesons is then determined solely by the behavior of the meson propagators. \begin{figure}[t] \begin{center} \epsfig{figure=./fig3.eps,height=7.0cm,width=7.0cm} \end{center} \caption{ The meson transition form factor $f^{VPP}(p_1,p_2)$ obtained from \Eqs{LamVPP:Quark} (circles) and the numerical fit (solid curve) of \Eq{VPPFit} used in the calculations of vector meson self energies and EM form factors. } \label{Fig:fVPP} \end{figure} \begin{figure}[t] \begin{center} \epsfig{figure=./fig4.eps,height=7.0cm,width=7.0cm} \end{center} \caption{ The meson transition form factor $f^{VVP}$ obtained from \Eqs{LamVVP:Quark} (circles) and the numerical fit (solid curve) of \Eq{VVPFit} used in the calculations of vector meson self energies. } \label{Fig:fVVP} \end{figure} Rather than directly use the numerical values of the meson transition form factors as calculated from \Eqs{LamVPP:Quark} and (\ref{LamVVP:Quark}), it is found that the resulting form factors can be parametrized in terms of the following functions: \begin{eqnarray} f^{VPP}(k\+\frac{q}{2},\-k\+\frac{q}{2}) &=& \exp\left(-\frac{k^2 \+ m_{P}^2 \- \sfrac{1}{4}m_{V}^2} {b_{VPP}^2(q^2)}\right) , \label{fVPPForm} \\ f^{VVP}(k\+\frac{q}{2},\-k\+\frac{q}{2}) &=& \left( \frac{1 \- m_{P}^2/b_{VPP}^2(q^2)}{1 \+ k^2/b_{VPP}^2(q^2)} \right)^2 \label{fVVPForm}, \end{eqnarray} where $p_1 = \sfrac{1}{2} q + k$ and $p_2 = \sfrac{1}{2} q - k$ in \Eq{fVPPForm} and \Eq{fVVPForm}. The parameters are chosen by fitting these forms to the numerical results on the domain $k^2 \in$ (0, 1)~GeV$^2$ for a range of external vector-meson four momenta $q^2 \in$ ($-1$, $+1$)~GeV$^2$. The $q^2$-dependence of the resulting parameters is well described in terms of the linear forms: \begin{eqnarray} b_{VPP}(q^2) &=& b_{VPP}^0 + b_{VPP}^1 \; q^2 , \nonumber \\ b_{VPP}^0 &=& 0.855 \;\;{\rm GeV} , \nonumber \\ b_{VPP}^1 &=& 0.0693 \;\;{\rm GeV}^{-1}, \label{VPPFit} \end{eqnarray} and \begin{eqnarray} b_{VVP}(q^2) &=& b_{VVP}^0 + b_{VVP}^1 \; q^2 , \nonumber \\ b_{VVP}^0 &=& 1.098 \;\;{\rm GeV} , \nonumber \\ b_{VVP}^1 &=& 0.159 \;\;{\rm GeV}^{-1} \label{VVPFit} . \end{eqnarray} The numerical results (denoted by circles) and the fits (solid curves) are shown in Figs.~\ref{Fig:fVPP} and \ref{Fig:fVVP} for $q^2 = -m_{\rho}^2$ for the form factors $f^{VPP}(p_1,p_2)$ and $f^{VVP}(p_1,p_2)$, respectively. The similarity of the two curves suggests that both could have just as easily been fitted to either exponential or dipole forms. The choice of one form over the other is arbitrary, and does not affect the results obtained herein. \section{Results and discussion} \label{Sec:Results} \subsection{mass dependence} \label{Sec:ResultsMassDep} \begin{figure}[t] \begin{center} \epsfig{figure=./fig5.eps,height=7.0cm} \end{center} \caption{ The real (solid curves) and imaginary (dashed curves) parts of the vector-meson self energy due to intermediate $\pi\pi$ and $KK$ meson states. The vertical dashed line denotes the on-mass-shell point $q^2 = - m_{\rho}^2$ for the $\rho$ meson and the vertical solid line denotes $q^2 = 0$. } \label{Fig:VPP} \end{figure} \Fig{Fig:VPP} shows the real (solid curves) and imaginary (dashed curves) parts of the contributions to the vector self energy due to $\pi\pi$ and $KK$ loops as a function of the four-momentum squared of the vector meson $q^2$. These curves are obtained by calculating $\Pi^{PP}(q^2,m_P,m_P)$ with $P=$ $\pi$ or $K$ in \Eqs{PiPP} and (\ref{FPP}) with the assumption of exact SU(3)-flavor symmetry; i.e., $g_{\rho\pi\pi} = g_{\rho K K} =$ 6.03 and using the parametrization of $f^{VPP}(p_1,p_2)$ given by \Eq{fVPPForm}. For spacelike values of the vector-meson momenta $q^2 > 0$ (to the right of the solid vertical line in \Fig{Fig:VPP}), the self energy $\Pi^{PP}(q^2,m_P,m_P)$ is purely real and approaches zero monotonically from below as $q^2 \rightarrow + \infty$. The threshold for the decay $\rho \rightarrow\pi\pi$ is at $q^2 = - 4 m_{\pi}^2$ (just left of the solid vertical line in \Fig{Fig:VPP}). Here, the imaginary part of the self energy $\Pi^{PP}(q^2,m_{\pi},m_{\pi})$ becomes non-zero. This abrupt change of the imaginary part of the self energy at the two-pion threshold leads to a discontinuity in the {\em second derivative} of the real part of the self energy. This causes the real part of the self energy to rapidly turn over at $q^2 \approx$ $-0.3$~GeV$^2$ and start to approach zero as $q^2$ becomes more negative (timelike). The same behavior is observed in the two-$K$-meson self energy $\Pi^{PP}(q^2,m_{K},m_{K})$, but much deeper in the timelike region since the two-kaon decay threshold is at $q^2 = -4 m^2_{K} \approx$ $-1$~GeV$^2$. The discontinuity in the second derivative of ${\rm Re}\Pi^{PP}(q^2,m_{P},m_{P})$ at $q^2 = -4 m_{P}^2$ is a general feature of an amplitude near the threshold of a decay into a two-meson intermediate state with $L=1$ relative angular momentum. In general, the dependence of the imaginary part of the self energy on the center-of-momentum energy $\sqrt{s}$ near a decay threshold is proportional to $\sqrt{s}^{(1 + 2 L)}$, where $L$ is the relative angular momentum of the two outgoing decay products. The centrifugal barrier due to the relative angular momentum $L$ tends to soften the behavior of the imaginary part of the amplitude at threshold. In the present case of $\rho\rightarrow\pi\pi$, the pions have relative $L=1$ angular momentum and hence the second derivative of ${\rm Re}\Pi^{PP}(q^2,m_{P},m_{P})$ is discontinuous at $q^2 = -4 m_{P}^2$. Had this two-body decay process proceeded in the $s$-channel with $L=0$, the {\em first} derivative of the self energy would have been observed as being discontinuous at threshold $q^2 = -4 m_{P}^2$. \begin{figure}[t] \begin{center} \epsfig{figure=./fig6.eps,height=7.0cm} \end{center} \caption{ The real (solid curves) and imaginary (dashed curves) parts of the vector-meson self energy due to intermediate $\omega\pi$ and $K^* K$ meson states. The vertical dashed line denotes the on-mass-shell point $q^2 = - m_{\rho}^2$ for the $\rho$ meson and the vertical solid line denotes $q^2 = 0$. } \label{Fig:VVP} \end{figure} In \Fig{Fig:VVP} the real (solid curves) and imaginary (dashed curves) parts of the vector-meson self energy due to vector-pseudoscalar intermediate state $\Pi^{VP}(q^2,m_V,m_P)$ is shown as a function of $q^2$. These curves are obtained by calculating $\Pi^{VP}(q^2,m_V,m_P)$ with $P=\pi$ and $V=\rho$ or $P=K$ and $V=K^*$ in \Eqs{PiVP} and (\ref{FVP}) with the assumption of exact SU(3)-flavor symmetry; i.e., $g_{\rho\omega\pi} = g_{\rho K^* K} =$ 10.81 and using the parametrization of $f^{VVP}(p_1,p_2)$ given by \Eq{fVVPForm}. Although not apparent from \Fig{Fig:VVP}, when the vector-meson four momentum becomes large and spacelike, $q^2 \rightarrow +\infty$, the real part of the self energy $\Pi^{VP}(q^2,m_V,m_P)$ approaches zero from above. The fact that the decay threshold for $\rho\rightarrow\omega\pi$ is observed at $q^2 = -(m_{\omega}+m_{\pi})^2 \approx$ -0.84~GeV$^2$, so far above the on-mass-shell value of $q^2 = -m_{\rho}^2$ means that very little structure for the real part of $\Pi^{VP}(q^2,m_{\omega},m_{\pi})$ is observed for the range of $q^2$ plotted in \Fig{Fig:VVP}. The lack of structure in the real part of $\Pi^{VP}(q^2,m_{K^*},m_{K})$ for the $K^*K$ channel is even more apparent, owing to the fact that the threshold for $\rho \rightarrow K^* K$ is found at $q^2 = -(m_{K^*}+m_{K})^2 \approx$ -1.9~GeV$^2$ (not shown in \Fig{Fig:VVP}). As discussed above for $\Pi^{PP}(q^2,m_P,m_P)$, the order in which the discontinuity in derivatives of the real part of the self energy appears is determined by the relative angular momentum $L$ of the out-going decay products. It is straight-forward to show that the decay of a vector meson with $J^P = 1^-$ into a vector and pseudoscalar meson with $J^P = 0^-$ must be in an orbital angular momentum $L = 1$ state. Hence, the discontinuity is expected to appear in the second derivative of ${\rm Re}\Pi^{VP}(q^2,m_V,m_P)$ at $q^2 =-(m_{\omega}+m_{\pi})^2$. Numerical differentiation of the results shown in \Fig{Fig:VVP} confirm the presence of this discontinuity. One observes that $\Pi^{VP}(q^2,m_V,m_P)$ passes through zero at $q^2 = 0$, a feature that is not observed for $\Pi^{PP}(q^2,m_P,m_P)$. This zero is a result of the structure of the $VVP$ meson transition amplitude $\Lambda^{ij}_{\mu\nu}(p_1,p_2)$ given in \Eq{VVPvertex}, which is required by Lorentz covariance. The requirement that this amplitude be proportional to \begin{equation} \epsilon_{\mu\nu\alpha\beta} \; p_{1\alpha}\; p_{2\beta} = \epsilon_{\mu\nu\alpha\beta} \; q_{\alpha} \; p_{1\beta} , \end{equation} leads to a vector-meson self energy $\Pi_{\mu\nu}^{VP}(q^2,m_V,m_P)$ that is necessarily transverse to $q_{\mu}$ for all values of $q^2$ and hence is of the form given in \Eq{Pimunu}. It follows that the vector-meson self energy in this channel vanishes at $q^2 = 0$. Because the real part of $\Pi_{\mu\nu}^{VP}(q^2,m_V,m_P)$ goes through zero at $q^2 = 0$, approaches zero in the deep spacelike region and is infinitely differentiable below the breakup threshold at $q^2 = -(m_V+m_P)^2$, it is a small and slowly varying function for $q^2 \geq -m_{\rho}^2$. Therefore, the $\omega\pi^0$ intermediate state is expected to contribute very little to the $\rho$ meson self energy (from \Fig{Fig:VVP}, it is clear that the $K^* K$ state contributes even less). However, this channel is important for the self energy of the $\omega$ meson, owing to the fact that there are three different $\rho\pi$ channels that contribute to the $\omega$-meson self energy; they are $\rho^+\pi^-$, $\rho^-\pi^+$, and $\rho^0\pi^0$. This leads to the overall factor of three in front of $\Pi_{\mu\nu}^{VP}(q^2,m_{\rho},m_{\pi})$ in \Eq{PiSum:Omega}, which makes the $\rho\pi$ channel as important for the $\omega$-meson self energy as the $\pi\pi$ channel is for the $\rho$-meson self energy (see \Tab{Tab:Shifts}). \begin{figure}[t] \begin{center} \epsfig{figure=./fig7.eps,height=7.0cm} \end{center} \caption{ The dependence of the real part of the vector-meson self energy on the mass $m_1$ of the intermediate state mesons. The dashed curve shows the dependence of the vector-pseudoscalar channel $\Pi^{VP}(-m_{\rho}^2,m_\rho,m_1)$ when the mass $m_1$ of the pseudoscalar meson is varied. The solid curve shows the dependence of the two-pseudoscalar channel $\Pi^{PP}(-m_{\rho}^2,m_{\pi},m_1)$ when the mass $m_1$ of {\em one} of the pseudoscalar mesons is varied. The short-dashed curve shows the dependence of the two-pseudoscalar channel $\Pi^{PP}(-m_{\rho}^2,m_1,m_1)$ when the mass $m_1$ of {\em both} pseudoscalar mesons is varied. } \label{Fig:m1dep} \end{figure} A single coupling constant and meson-transition form factors has been used for each of the two types of intermediate states ($PP$ and $VP$) considered herein. The result is that the only difference between the various channels within a particular type of meson loop is the masses of the intermediate mesons involved. A comparison of the self energies resulting from the $KK$ and $K^*K$ channels with $\pi\pi$ and $\omega\pi$ channels, respectively, suggests that states with heavier mesons contribute less than states of lighter mesons. For example, for all $q^2 \geq - m_{\rho}^2$, the value of $\Pi^{PP}(q^2,m_K,m_K)$ is less than half the value of $\Pi^{PP}(q^2,m_{\pi},m_{\pi})$. The same is true when $\Pi^{VP}(q^2,m_{K^*},m_K)$ is compared with $\Pi^{PP}(q^2,m_{\omega},m_{\pi})$ in \Fig{Fig:VVP}. Within the present model, the dependence of the vector-meson self energy on the mass of the intermediate mesons can be explored. This is done by letting one or both of the masses $m_P$ in \Eqs{PiPP} and (\ref{PiVP}) vary while keeping $q^2 = -m_{\rho}^2$. The solid curve in \Fig{Fig:m1dep} is the result of a calculation of the real part of the self energy $\Pi^{PP}(-m_{\rho}^2,m_{\pi},m_1)$ in which one of the pseudoscalar-meson propagators in \Eq{PiPP} has mass $m_{\pi}$ and the other propagator has a mass $m_1$ which is varied from 0 to 1500~MeV. It is clear from \Fig{Fig:m1dep} that the magnitude of the self energy depends critically on the mass $m_1$. In particular, for $m_1 = m_{\rho}$, the real part of the self energy $\Pi^{PP}(-m_{\rho}^2,m_{\pi},m_1)$ has already decreased by a factor of two from its value for $m_1 = m_{\pi}$. The short-dashed curve in \Fig{Fig:m1dep} is the self energy $\Pi^{PP}(-m_{\rho}^2,m_1,m_1)$ obtained with the masses of the two pseudoscalar mesons are taken to be {\em equal} and varied together. Increasing both masses of the intermediate mesons at the same time causes the self energy to decrease very rapidly. It was argued in \Ref{LC} that lattice calculations which linearly extrapolate the $\rho$ mass as a function of $m_\pi$ should receive negligible corrections from the two-pion intermediate state. The calculations described here are in agreement with this conclusion, but as can be seen from the short-dashed curve in \Fig{Fig:m1dep}, the dependence of the real part of the vector self-energy in the region of the physical pion mass is not smooth, showing the effects of the threshold crossed at $4\,m_\pi^2$. This rapid dependence on the mass of the intermediate states suggests that such extrapolations must be made with caution. The dashed curve in \Fig{Fig:m1dep} shows the dependence of the self energy $\Pi^{VP}(-m_{\rho}^2,m_{\omega},m_1)$ when the mass of the intermediate vector meson is $m_V$ and the mass of the intermediate pseudoscalar meson $m_1$ is varied. Because of the different Lorentz-covariant structure of the vector-vector-pseudoscalar transition amplitude $\Lambda_{\mu\nu}^{ij}(p_1,p_2)$ given in \Eq{VVPvertex}, and that the threshold for $\rho \rightarrow \omega \pi$ has not been reached even for $m_1 = 0$, a much weaker dependence on the mass $m_1$ is observed compared to that of $\Pi^{PP}(-m_{\rho}^2,m_1,m_2)$. Nonetheless, it may be sufficiently accurate to neglect pseudoscalar-vector intermediate states with masses $m_1 \geq m_{\rho}$ when calculating the self energies, since these are smaller by at least a factor of two than the contribution from the $\omega\pi$ intermediate state. It is interesting to note that the curves in \Fig{Fig:m1dep} all exhibit an extremum for values of $m_1$ for which the sum of the intermediate-mesons masses $(m_1 + m_2)$ is within the range $\sfrac{1}{2} m_{\rho} \leq (m_1 + m_2) \leq m_{\rho}$. This is not coincidental, since $m_{\rho}$ is the value of the center-of-momentum energy at the decay threshold. For values of $m_1 > m_{\rho} - m_2$ the real part of the amplitude is a monotonic, decaying function of $m_1$. At threshold $m_1 = m_{\rho} - m_2$, discontinuities in the second derivative of the real part of the self energy cause the rapid turnover observed in \Fig{Fig:m1dep}, so that the extremum of ${\rm Re }\Pi(-m^2_{\rho},m_1,m_2)$ occurs for a value of $m_1$ slightly below the threshold value $m_1 = m_{\rho} - m_2$. This suggests that in calculations of self energies for a particle of mass $M$, it may be sufficiently accurate to consider only those states for which the sum of the intermediate hadron masses $m_1 + m_2 + \cdots$ lie on the range $\sfrac{1}{2} M \leq \sum_{i} m_i \leq M$. This implies greatly simplified calculations of mixings between states where it is only necessary to consider the contributions from multiple-hadron states within this range of total masses. This is a general feature of such calculations and not a specific result of the particular channels considered herein, since it arises independently of the particular form and scales used to model the meson transition form factors $f^{VPP}(p_1,p_2)$ and $f^{VVP}(p_1,p_2)$. \subsection{vector-meson mass shifts} \label{Sec:ResultsMassShifts} \begin{table}[t] \caption {Mass shifts to $\rho$ and $\omega$ mesons arising from several two-meson intermediate states. Each channel provides less than a $10\%$ correction to the total mass of the vector meson. The sum of these shifts, which is not observable, is given in the bottom row. The {\em difference} between these sums is $m_{\omega} - m_{\rho} =$ 24.8~MeV, which can be compared to the experimental value of $12\pm 1$ MeV. } \begin{tabular}{c|rr} channel & $\Delta m_{\rho}$ (MeV)& $\Delta m_{\omega}$ (MeV)\\ \hline $\pi\pi$ & -69.8 & \\ $K \bar{K} $ & -14.8 & -14.8 \\ $\omega\pi$ & -22.5 & \\ $\rho\pi$ & & -67.5 \\ $\omega\eta$ & & -13.1 \\ $\rho\eta$ & -13.1 & \\ $K^* K $ & -11.9 & -11.9 \\ \hline sum &-132.1 &-107.3 \\ \end{tabular} \label{Tab:Shifts} \end{table} Results for the mass shifts due to the two-meson loops considered in this study are summarized in \Tab{Tab:Shifts}. It should be noted that only the difference between the total mass shifts of the $\rho$ and $\omega$ mesons is physically meaningful. The absolute size of each term does, however, give some indication of the importance of the dressing by two-meson loops. The largest mass shifts are that of the $\rho$ mass due to the two-pion loop, and that of the $\omega$ due to the $\rho\pi$ loop, which are about $-10\%$ of the bare mass. Note that the $\rho$ mass shift due to the two-pion loop is roughly $-70$ MeV, while the two-pion $\rho$ mass shift in \Ref{LC} was between $-10$ and $-20$ MeV. The difference is due to an additional contribution to the vector self-energy which arises from a $\pi\pi\rho\rho$ contact term. This term provides a contribution to the vector-meson self energy $\Pi_{\mu\nu}(q)$ that is independent of $q$. In \Ref{LC}, this term is added to provide current conservation for the vector $\rho$; i.e., that $\Pi^{PP}(q^2=0)=0$. Since one can show that this term contributes equally to the self energies of the $\rho$ and $\omega$ mesons, it cannot contribute to their mass splitting and so has not been included here. To allow for the direct comparison with \Ref{LC}, the effects of including this contact term into the present study can be reproduced by subtracting $\Pi(q^2=0)$ from the self energy $\Pi(q^2)$. This procedure ensures that the condition obtained in \Ref{LC} that $\Pi^{PP}(q^2=0)=0$ is reproduced. The resulting mass shift due to the two-pion loop obtained from \Eq{massShifts} would then be approximately \begin{eqnarray} \delta m_{\pi\pi} & \approx & {\rm Re}\left[ \frac{ \Pi^{PP}(\-m_\rho^2,m_\pi,m_\pi) \- \Pi^{PP}(0,m_\pi,m_\pi) }{2 \; m_0}\right] \nonumber\\ & \approx & - 9.2\ {\rm MeV}. \end{eqnarray} This mass shift for the $\rho$ meson is similar to that of \Ref{LC}. As pointed out in \Ref{GI}, the strange-meson and vector-$\eta$ intermediate states cannot contribute to the mass splitting; they are included here to illustrate that as the masses of the intermediate state mesons increase, the resulting shifts decrease rapidly. For example, the shift in the $\rho$ mass due to the $K\bar{K}$ intermediate state is about one fifth of that due to two pions. Similarly, the shift in the $\omega$ mass due to the $\rho\eta$ intermediate state is roughly half the size of that due to $\omega\pi$. The result of adding the shifts due to the various intermediate states is that they largely cancel, as in \Ref{GI}, to give a mass splitting of roughly 25 MeV, which is of the same order (on the scale of the bare mass) as the experimental value of $12\pm 1$ MeV. Of course, a complete calculation would have to consider all possible multiple-hadron intermediate states. It is encouraging that, within this model, a reasonable value of the mass splitting is achieved using only a handful of meson loops, and that the individual contributions due to additional loops are expected to be small compared to the dominant contributions evaluated here. \subsection{time orderings} \label{Sec:ResultsOrders} \begin{figure}[t] \begin{center} \epsfig{figure=./figratios.eps,height=7.0cm,width=7.0cm} \end{center} \caption{ Ratios of the forward-forward (dashed curve), backward-backward (solid curve) and forward-backward or backward-forward (short-dashed long-dashed curve) time orderings of the real part of the two-pion loop integral to the sum of all four, as a function of the vector-meson mass squared.} \label{Fig:Ratios} \end{figure} Figure~\ref{Fig:Ratios} shows the ratio of the forward-forward $\Pi^{[++]}(q^2)$, backward-backward $\Pi^{[--]}(q^2)$, and forward-backward $\Pi^{[+-]}(q^2)$ [equal to backward-forward $\Pi^{[-+]}(q^2)$] time orderings of the real part of the two-pion loop integral to their sum $\Pi^{\pi\pi}(q^2)$, as a function of the vector-meson mass squared $q^2$. As expected from the analysis in Sec.~\ref{Sec:Orders}, at $q^2=0$ the forward-forward time ordering $\Pi^{[++]}(q^2)$ is equal to the backward-backward time ordering $\Pi^{[--]}(q^2)$. It is also simple to show [by a change of variables from $k_4$ to $-k_4$ in Eq.~\ref{TOPiPP}] that $\Pi^{[+-]}(q^2)=\Pi^{[-+]}(q^2)$ at all values of $q^2$. Although not shown here, an examination of the derivative of $\Pi^{[++]}(q^2)$ indicates a discontinuity at $q^2=4m_\pi^2$ indicating the onset of a branch cut in $\Pi^{\pi\pi}(q^2)$ due to $\Pi^{[++]}(q^2)$, which verifies the identification of this time ordering with the propagation of two forward-propagating pions. In Fig.~\ref{Fig:Ratios}, one observes that at $q^2 = - m_{\rho}^2$ the forward-forward time ordering is dominant and the backward-backward contribution is only a few percent of the total. This realizes the expected strong suppression of the backward-backward time ordering, due to the vertices and the large energy denominators for the $\rho\rho\pi\pi$ intermediate state. Note also that both the forward-forward and backward-backward contributions have the same sign, so that the two-pion loop contribution to the $\rho$ self energy is not reduced by a cancellation between these two time orderings. This establishes that, in the present quantum field-theoretic framework, the dominant contribution to the $\rho$ self energy is from two forward-propagating pions, and that the difference between the mass shifts obtained in this approach and those of the quantum-mechanical approach of Ref.~\cite{GI} is not due to the presence of additional time orderings in a Lorentz covariant framework. As discussed in Sec.~\ref{TOPiPP}, the forward-backward and backward-forward time orderings are not zero in the present framework; Fig.~\ref{Fig:Ratios} shows that they also make negligible contributions to $\Pi^{\pi\pi}(q^2)$ at $q^2=-m_\rho^2$. \subsection{vector-meson electromagnetic form factors} \label{Sec:ResultsEMffs} \begin{figure}[t] \begin{center} \epsfig{figure=./fig8.eps,height=7.0cm,width=7.0cm} \end{center} \caption{ The $\rho$-meson electric-monopole form factor. The dashed curve is the result $G_{E}^{\bar{q}q}(q^2)$ from \Ref{Hawes} that rises from quark-antiquark substructure of $\rho$ meson. The solid curve is the real part of the form factor $G_{E}(q^2)$ that results when the effects of pion loops from \Eq{Gpipik2} are added to the results from \Ref{Hawes}. } \label{Fig:Ge} \end{figure} The final application of vector-meson dressing by two-meson loops that is considered here is the correction to the EM form factors and charge radius of the $\rho$ meson. In \Sec{Sec:EMFF}, an expression for the pion contribution to the EM form factor $G_{i}^{\pi\pi}(Q^2)$ was obtained in terms of the same meson transition form factor $f^{VPP}(p_1,p_2)$, given in \Eq{fVPPForm} with \Eq{VPPFit}, that was used to obtain the $\pi\pi$ and $K\bar{K}$ contributions to the $\rho$-meson self energy. In the present calculation of $G^{\pi\pi}_E(Q^2)$, the final term in \Eq{Gpipik2} (arising from the surface contribution of the integration by parts) is small and varies slowly with $Q^2 \approx 0$, hence it is safe to neglect this term for the determination of the charge radius. The resulting $G_{E}^{\pi\pi}(Q^2)$ is well described for spacelike values of the photon momentum squared $0 < Q^2 < $ 0.8~GeV$^2$ by the form \begin{equation} G_{E}^{\pi\pi}(Q^2) \approx A \; e^{-B Q^2 / m_{\rho}^2} \end{equation} with $B \approx$ 1.39 and the value of $A$ as determined from \Eq{deltaZpipi} is $A = \delta Z_{\pi\pi} = $ 0.134. The form factor $G_{E}^{\pi\pi}(Q^2)$ is added to the quark-antiquark contribution $G_{E}^{q\bar{q}}(Q^2)$ obtained by \Ref{Hawes} according to \Eq{Ge}. The resulting form factor $G_{E}^{\pi\pi+q\bar{q}}(Q^2)$ is plotted as a solid curve in \Fig{Fig:Ge}. It falls more rapidly with the photon momentum $Q^2$ than does the form factor $G_{E}^{q\bar{q}}(Q^2)$ (dashed curve) indicating that the inclusion of the $\pi\pi$ intermediate state into the $\rho$ meson self energy has {\em increased} the charge radius of the $\rho$ meson. The charge radius of the $\rho$ meson due to its quark-antiquark substructure and its mixing with the two-pion state obtained using \Eq{RadiusDef} is \begin{equation} {\langle r^2_{\bar{q}q + \pi\pi} \rangle }^{\sfrac{1}{2}} = 0.67 \;\;{\rm fm}. \end{equation} This value may be compared to the value obtained in \Ref{Hawes}, \begin{equation} {\langle r^2_{\bar{q}q} \rangle }^{\sfrac{1}{2}} = 0.61 \;\;{\rm fm} , \end{equation} which results from only the quark-antiquark substructure of the $\rho$ meson. One observes that including pion loops increases the charge radius of the $\rho$ meson by approximately $10\%$. This is consistent with the increase of less than $15\%$ observed in a similar study of the pion-loop contribution to the $\pi$ meson charge radius \cite{Bender}. Although, the short lifetime of the $\rho$ meson excludes the possibility of directly measuring the EM charge radius of the $\rho$ meson at present, one can define a ``diffractive radius'' from the $t$ dependence of diffractive $\rho$-meson electroproduction cross sections \cite{MAP}. As with the EM charge radius, the diffractive radius receives contributions from its quark substructure as well as from mixings with multiple-hadron states. However, an important distinction between the charge radius and the diffractive radius is that since diffraction arises from a strong interaction, electromagnetically neutral particles, such as the $\omega$ and $\rho^0$ mesons, will have diffractive radii even though they have no well-defined charge radius\footnote { To be electroproduced diffractively, the vector meson must be able to mix with the photon, and so must be neutral (e.g., $\omega$, $\rho^0$, $\phi$, $J/\psi$). Therefore, {\em only} for neutral vector mesons can one define a diffractive radius. }. The results of this study suggest that the most significant contribution from two-meson loops to the diffractive radius of the $\rho$ meson would be due to the two-pion loop. However, $G$-parity forbids the $\omega$ meson from mixing with the two pions, so that the diffractive radius of the $\omega$ meson would be unaffected by the two-pion loop. Therefore, the diffractive radius of the $\omega$ meson is expected to be {\em smaller} than that of the $\rho$ meson. An observed difference in the $t$-dependence of diffractive electroproduction of $\rho$ and $\omega$ mesons would be indicative of their differing diffractive radii, and such a measurement could provide a means to estimate the contributions to these radii from pion loops. \section{Conclusions} \label{Sec:Concl} It is shown that, in a covariant model based on studies of the Schwinger-Dyson equation of QCD which assumes exact SU(3)-flavor symmetry, the contributions to the self energies of the $\rho$ and $\omega$ due to several pseudoscalar-pseudoscalar and pseudoscalar-vector meson loops are at most $10\%$ of the bare mass. The result for the mass shift of the $\rho$ meson due to the two-pion loop is in agreement with a previous calculation of this quantity using an effective chiral Lagrangian approach. Such contributions are found to decrease rapidly as the mass of the intermediate mesons increases beyond $m_{\rho}/2$. The mass shifts due to several two-meson intermediate states are compared with those from an extensive study within a nonrelativistic framework of the $\rho$-$\omega$ mass splitting, and are found to be smaller, especially for intermediate states involving higher-mass mesons. A mass splitting of $m_{\omega} - m_{\rho}\approx 25$~MeV is found from the $\pi\pi$, $K\bar{K}$, $\omega\pi$, $\rho\pi$, $\omega\eta$, $\rho\eta$ and $K^* K$ channels. A complete calculation which evaluates contributions from all possible multiple hadron intermediate states and breaks SU(3)-flavor symmetry is beyond the scope of the present work. However, these results suggest that such a calculation should exhibit rapid convergence as the number of two-meson intermediate states is increased by including states with higher masses. This implies that inclusion of two-meson loops into the vector-meson self-energy yields a small correction to the predominant valence quark-antiquark structure of the vector meson. The part of the vector-meson self energy which corresponds to two pions propagating forward in time has been shown to dominate a Lorentz-covariant calculation of the two-pion loop contribution. This implies that the additional time-orderings of this loop, necessary to maintain Lorentz covariance, are not responsible for the reduced size of the mass shifts found in the present framework when compared to quantum-mechanical treatments. As a test of the self-consistency of this calculation, the contribution of the two-pion loop to the $\rho$ meson EM form factor is evaluated, and is shown to provide a modest increase of about $10\%$ to the charge radius of the $\rho$. Although this charge radius is difficult to observe, it may be possible to observe the effects of such an increase by examining the difference in the ``diffractive radii'' of the neutral $\rho$ and $\omega$ found from the $t$-dependence of diffractive electroproduction cross sections. As the two-pion loop cannot contribute to the $\omega$, it may be possible to observe its effects on the $\rho$ by measuring a difference in these diffractive radii. \section{Acknowledgments} The authors wish to thank A.~Szczepaniac for illuminating discussions. This work is supported by the U.S. Department of Energy under contracts DE-AC05-84ER40150, DE-FG05-92ER40750 and DE-FG02-87ER40365, and the Florida State University Supercomputer Computations Research Institute which is partially funded by the Department of Energy through contract DE-FC05-85ER25000.
\section{Introduction} \label{sec:1} Transport processes play a crucial role in many natural phenomena. Among the many examples, we just mention the particle transport in geophysical flows which is of obvious interest for atmospheric and oceanic issues. The most natural framework for investigating such phenomena is to adopt a Lagrangian viewpoint in which the particles are advected by a given Eulerian velocity field $\bbox{u}(\bbox{x},t)$ according to the differential equation \begin{equation} {d \bbox{x} \over d t} = \bbox{u}(\bbox{x},t) = \bbox{v}(t)\,, \label{eq:1.1} \end{equation} where, by definition, $\bbox{v}(t)$ is the Lagrangian particle velocity. Despite the apparent simplicity of (\ref{eq:1.1}), the problem of connecting the Eulerian properties of $\bbox{u}$ to the Lagrangian properties of the trajectories $\bbox{x}(t)$ is a very difficult task. In the last $20-30$ years the scenario has become even more complex by the recognition of the ubiquity of Lagrangian chaos (chaotic advection). Even very simple Eulerian fields can generate very complex Lagrangian trajectories which are practically indistinguishable from those obtained in a complex, turbulent, flow \cite{H66,Licht,Ottino,lagran,zav,CM93}. Despite these difficulties, the study of the relative dispersion of two particles can give some insight on the link between Eulerian and Lagrangian properties at different length-scales. Indeed, the evolution of the separation $\bbox{R}(t)=\bbox{x}^{(2)}(t)-\bbox{x}^{(1)}(t)$ between two tracers is given by \begin{equation} {d \bbox{R} \over d t} = \bbox{v}^{(2)}(t)-\bbox{v}^{(1)}(t)= \bbox{u}(\bbox{x}^{(1)}(t)+\bbox{R}(t),t)- \bbox{u}(\bbox{x}^{(1)}(t),t) \label{eq:1.5} \end{equation} and thus depends on the velocity difference on scale $\bbox{R}$. It is obvious from (\ref{eq:1.5}) that Eulerian velocity components of typical scale much larger than $\bbox{R}$ will not contribute to the evolution of $\bbox{R}$. Since, in incompressible flows, separation $\bbox{R}$ typically grows in time \cite{Cocke69,Orszag70} we have the nice situation in which from the evolution of the relative separation we can, in principle, extract the contributions of all the components of the velocity field. For these reasons, in this paper we prefer to study relative dispersion instead of absolute dispersion. For spatially infinite cases, without mean drift there is no difference; for closed basins the relative dispersion is, for many aspects, more interesting than the absolute one, which is dominated by the sweeping induced by large scale flow. There are very few general results on the link between Eulerian and Lagrangian properties and only for asymptotic behaviors. Let us suppose that the Eulerian velocity field is characterized by two typical length-scales: the (small) scale $l_u$ below which the velocity is smooth, and a (large) scale $L_0$ representing the size of the largest structures present in the flow. Of course, in most non turbulent flows it will turn out that $\l_u \sim L_0$. At very small separations $R \ll \l_u$ we have that the velocity difference in (\ref{eq:1.5}) can be reasonably approximated by a linear expansion in $R$, which in most time-dependent flows leads to an exponential growth of the separation of initially close particles, a phenomenon known as Lagrangian chaos \begin{equation} \langle \ln R(t) \rangle \simeq \ln R(0) + \lambda t \label{eq:1.6} \end{equation} (the average is taken over many couples with initial separation $R(0)$). The coefficient $\lambda$ is the Lagrangian Lyapunov exponent of the system \cite{Licht}. The rigorous definition of the Lyapunov exponent imposes to take the two limits $R(0) \to 0$ and then $t \to \infty$: in physical terms these limits amount to the requirement that the separation has not to exceed the scale $\l_u$ but for very large times. This is a very strict condition, rarely accomplished in real flows, rendering often infeasible the experimental observation of the behavior (\ref{eq:1.6}). On the opposite limit, for very long times and for separations $R \gg L_0$, the two trajectories $\bbox{x}^{(1)}(t)$ and $\bbox{x}^{(2)}(t)$ feel two velocities which can practically be considered as uncorrelated. We thus expect normal diffusion, i.e. \begin{equation} \langle R^2(t) \rangle \simeq 2 D\, t\,. \label{eq:1.7} \end{equation} Also in this case it is necessary to remark that the asymptotic behavior (\ref{eq:1.7}) cannot be attained in many realistic situations, the most common of which is the presence of boundaries at a scale comparable with $L_0$. In absence of boundaries it is possible to formulate sufficient conditions on the nature of the Eulerian flow, under which normal diffusion (\ref{eq:1.7}) always takes place asymptotically \cite{MA89-AV95}. Between the two asymptotic regimes (\ref{eq:1.6}) and (\ref{eq:1.7}) the behavior of $R(t)$ depends on the particular flow. The study of the evolution of the relative dispersion in this crossover regime is very interesting and can give an insight on the Eulerian structure of the velocity field. To summarize, in all systems in which the characteristic length-scales are not sharply separated, it is not possible to describe dispersion in terms of asymptotic quantities. In such cases, different approaches are required. Let us mention some examples: the symbolic dynamics approach to the sub-diffusive behavior in a stochastic layer \cite{A} and to mixing in meandering jets \cite{B}; the study of tracer dynamics in open flows in terms of chaotic scattering \cite{C} and the exit time description for transport in semi-enclosed basins \cite{D} and open flows \cite{E}. The aim of the present paper is to discuss the use of an indicator -- the Finite Size Lyapunov Exponents (FSLE), originally introduced in the context of predictability problems \cite{ABCPV97} -- to study and characterize non-asymptotic transport in non-ideal systems, e.g. closed basins and systems in which the characteristic length-scales are not sharply separated. In section \ref{sec:2} we introduce the basic tools for the finite-scale analysis and we discuss their general properties. Section \ref{sec:3} is devoted to the evaluation of our method on some numerical examples. We shall see that even in apparently simple situations the use of finite scale analysis avoids possible misinterpretation of the results. In section \ref{sec:4} the method is applied to two physical problems: the analysis of experimental drifter data and the numerical study of relative dispersion in fully developed turbulence. Conclusions are presented in section \ref{sec:5}. The appendices report, for sake of self-consistency, some technical aspects. \section{Finite size diffusion coefficient} \label{sec:2} In order to introduce the finite size analysis for the dispersion problem let us start with a simple example. We consider a set of N particle pairs advected by a smooth (e.g. spatially periodic) velocity field with characteristic length $l_{u}$. Denoting with $R_i^{2}(t)$ the square separation of the $i$-th couple, we define \begin{equation} \langle R^{2}(t) \rangle= {1 \over N} \sum_{i=0}^N R_i^2 \, . \label{def:disprel} \end{equation} We assume that the Lagrangian motion is chaotic, thus we expect the following regimes to hold \begin{equation} \langle R^{2}(t) \rangle \simeq \left\{ \begin{array}{ll} R^{2}_{0}\exp(L(2)t) & \;\;\;\; {\mbox {if $\langle R^{2}(t)\rangle^{1/2} \ll l_{u}$}} \\ 2 D t & \;\;\;\; {\mbox {if $\langle R^{2}(t)\rangle^{1/2} \gg l_{u}$}} \end{array} \label{eq:regimiperR} \right. \,, \label{example1} \end{equation} where $L(2) \geq 2\lambda$ is the generalized Lyapunov exponent \cite{BPPV85,PV87}, $D$ is the diffusion coefficient and we assume that $R_i(0)=R_0$. An alternative method to characterize the dispersion properties is by introducing the ``doubling time\/'' $\tau(\delta)$ at scale $\delta$ as follows \cite{ABCCV97}: given a series of thresholds $\delta^{(n)}= r^{n} \delta^{(0)}$, one can measure the time $T_i(\delta^{(0)})$ it takes for the separation $R_i(t)$ to grow from $\delta^{(0)}$ to $\delta^{(1)}= r \delta^{(0)}$, and so on for $T_i(\delta^{(1)})\,,\;T_i(\delta^{(2)})\,,\ldots$ up to the largest considered scale. The $r$ factor may be any value $> \, 1$, properly chosen in order to have a good separation between the scales of motion, i.e. $r$ should be not too large. Strictly speaking, $\tau(\delta)$ is exactly the doubling time if $r\,=\,2$. Performing the doubling time experiments over the $N$ particle pairs, one defines the average doubling time $\tau(\delta)$ at scale $\delta$ as \begin{equation} \tau(\delta) = < T(\delta) >_e =\frac{1}{N} \sum_{i=1}^{N} T_{i}(\delta)\,. \label{def:taudelta} \end{equation} It is worth to note that the average (\ref{def:taudelta}) is different from the usual time average (see Appendix \ref{app:1}). Now we can define the Finite Size Lagrangian Lyapunov Exponent (see \cite{ABCPV97} for a detailed discussion) in terms of the average doubling time as \begin{equation} \lambda(\delta)=\frac{\ln r}{\tau(\delta)}\,, \end{equation} which quantifies the average rate of separation between two particles at a distance $\delta$. Let us remark that $\lambda(\delta)$ is independent of $r$, for $r$ close to $1$. For very small separations (i.e. $\delta \ll l_u$) one recovers the Lagrangian Lyapunov exponent $\lambda$, \begin{equation} \lambda=\lim_{\delta \rightarrow 0} \frac{1}{\tau(\delta)} \ln r\,. \label{def:liapfromtau} \end{equation} In this framework the finite size diffusion coefficient \cite{ABCCV97}, $D(\delta)$, dimensionally turns out to be \begin{equation} D(\delta)=\delta^{2}\lambda(\delta)\,. \label{def:fsd} \end{equation} Note the absence of the factor $2$, as one may expect from (\ref{eq:regimiperR}), in the denominator of $D(\delta)$; this is because $\tau(\delta)$ is a difference of times. For a standard diffusion process $D(\delta)$ approaches the diffusion coefficient $D$ (see eq. (\ref{eq:regimiperR})) in the limit of very large separations ($\delta \gg l_u$). This result stems from the scaling of the doubling times $\tau(\delta) \sim \delta^2$ for normal diffusion. Thus the finite size Lagrangian Lyapunov exponent $\lambda(\delta)$ behaves as follows: \begin{equation} \lambda(\delta) \sim \left\{ \begin{array}{ll} \lambda & \;\;\;\; {\mbox {if $\delta \ll l_{u}$}} \\ D/\delta^{2} & \;\;\;\; {\mbox {if $\delta \gg l_{u}$}} \end{array} \right. \,, \label{eq:regimipertau} \end{equation} One could naively conclude, matching the behaviors at $\delta \sim l_{u}$, that $D \sim \lambda l_{u}^{2}$. This is not always true, since one can have a rather large range for the crossover due to nontrivial correlations which can be present in the Lagrangian dynamics \cite{lagran}. One might wonder that the introduction of $\tau(\delta)$ is just another way to look at $\langle R^{2}(t)\rangle$. This is true only in limiting cases, when the different characteristic lengths are well separated and intermittency is weak. A similar idea of using times for the computation of the factor diffusion coefficient in nontrivial cases was developed in Ref. \cite{previouswork1,previouswork2,sabot}. If one wants to identify the physical mechanisms acting on a given spatial scale, the use of scale dependent quantities is more appropriate than time dependent ones. For instance, in presence of strong intermittency (which is indeed a rather usual situation) $R^{2}(t)$ as a function of $t$ can be very different in each realization. Typically one has (see figure~1a), different exponential growth rates for different realizations, producing a rather odd behavior of the average $\langle R^{2}(t)\rangle$ not due to any physical mechanisms. For instance in figure~1b we show the average $\langle R^{2}(t)\rangle$ versus time $t$; at large times one recovers the diffusive behavior but at intermediate times appears an ``anomalous'' diffusive regime which is only due to the superposition of exponential and diffusive contributions by different samples at the same time. On the other hand, by exploiting the tool of doubling times one has an unambiguous result (see figure~1c) \cite{ABCCV97}. An important physical problem where the behavior of $\tau(\delta)$ is essentially well understood is the relative dispersion in 3D fully developed turbulence. Here the smallest Eulerian scale $l_u$ is the Kolmogorov scale at which the flow becomes smooth. In the inertial range $l_u < R < L_0$ we expect the {\it Richardson law} to hold $\langle R^2(t) \rangle \sim t^3$; for separations larger than the integral scale $L_0$ we have normal diffusion. In terms of the finite size Lyapunov exponent we thus expect three different regimes: \begin{enumerate} \item $\lambda(\delta)=\lambda$ \,\,\,\, for $\delta \ll l_u$ \item $\lambda(\delta) \sim \delta^{-2/3}$ \,\,\,\, for $l_u \ll \delta \ll L_0$ \item $\lambda(\delta) \sim \delta^{-2}$ \,\,\,\, for $\delta \gg L_0$ \end{enumerate} We will see in section \ref{sec:4} than even for large Reynolds numbers, the characteristic lengths $l_u$ and $L_0$ are not sufficiently separated and the different scaling regimes for $\langle R^2(t) \rangle$ cannot be well detected. The fixed scale analysis in terms of $\lambda(\delta)$ for fully developed turbulence presents clear advantages with respect to the fixed time approach. \section{Numerical Results on simple flows} \label{sec:3} In this section we shall discuss some examples of applications of the above introduced indicator $\lambda(\delta)$ (or equivalently $D(\delta)$) for simple flows. The technical and numerical details of the finite size Lyapunov exponent computation are settled out in Appendix \ref{app:1}. In a generic case in addition to the two asymptotic regimes (\ref{eq:regimipertau}) discussed in section \ref{sec:2}, we expect another universal regime due to the presence of the boundary of given size $L_B$. For separations close to the saturation value $\delta_{max} \simeq L_B$ we expect the following behavior to hold for a broad class of systems \cite{ABCCV97}: \begin{equation} \lambda(\delta)=\frac{D(\delta)}{\delta^{2}} \propto \frac{(\delta_{max}-\delta)}{\delta} \,. \label{eq:nearbound} \end{equation} The proportionality constant is given by the second eigenvalue of the Perron-Frobenius operator which is related to the typical time of exponential relaxation of tracers' density to uniform distribution (see Appendix \ref{app:2}). \subsection{A model for transport in Rayleigh-B\'enard convection} \label{sec:3.1} The advection in two dimensional incompressible flows in absence of molecular diffusion is given by Hamiltonian equation of motion where the stream function, $\psi$, plays the role of the Hamiltonian: \begin{equation} \frac{dx}{dt}=\frac{\partial \psi}{\partial y}\,, \;\;\; \frac{dy}{dt}=-\frac{\partial \psi}{\partial x}\,. \label{eq:hamilton} \end{equation} If $\psi$ is time-dependent one typically has chaotic advection. As an example let us consider the time-periodic Rayleigh-B\'enard convection, which can be described by the following stream function \cite{gollub}: \begin{equation} \psi(x,y,t)=\frac{A}{k} \sin\left\{ k \left[ x+B \sin(\omega t)\right]\right\} W(y)\,, \label{eq:gollubinf} \end{equation} where $W(y)$ satisfies rigid boundary conditions on the surfaces $y=0$ and $y=a$ (we use $W(y)=\sin(\pi y/a)$). The two surfaces $y=a$ and $y=0$ are the top and bottom surfaces of the convection cell. The time dependent term $B\sin(\omega t)$ represents lateral oscillations of the roll pattern which mimic the even oscillatory instability \cite{gollub}. Concerning the analysis in terms of the finite size Lyapunov exponent one has that, if $\delta$ is much smaller than the domain size, $\lambda(\delta)=\lambda$. At larger values of $\delta$ we find standard diffusion $\lambda(\delta)=D/\delta^{2}$ with good quantitative agreement with the value of the diffusion coefficient evaluated by the standard technique, i.e. using $\langle R^{2}(t)\rangle$ as a function of time $t$. In order to study the effects of finite boundaries on the diffusion properties we confine the tracers' motion in a closed domain. This can be achieved by slightly modifying the stream function (\ref{eq:gollubinf}). We have modulated the oscillating term in such a way that for $|x|=L_B$ the amplitude of the oscillation is zero, i.e. $B \rightarrow B \sin(\pi x/L_B)$ with $L_B=2\,\pi n/k$ ($n$ is the number of convective cells). In this way the motion is confined in $x \in [-L_B,L_B]$. In figure~2 we show $\lambda(\delta)$ for two values of $L_B$. If $L_B$ is large enough one can distinguish the three regimes: exponential, diffusive and the saturation regime eq. (\ref{eq:nearbound}). Decreasing the size of the boundary $L_B$, the range of the diffusive regime decreases, while for small values of $L_B$, it disappears. \subsection{Point vortices in a Disk} \label{sec:3.2} We now consider a two-dimensional time-dependent flow generated by $M$ point vortices, with circulations $\Gamma_1,\dots,\Gamma_M$, in a disk of unit radius \cite{Aref}. The passive tracers are advected by the time dependent velocity field generated by the vortices and behave chaotically for any $M>2$. Let us note that in this case the scale separation is not imposed by hand, but depends on $M$ and on the energy of the vortex system \cite{BCF96}. Figure 3a shows the relative diffusion as a function of time in a system with $M=4$ vortices. Apparently there is an intermediate regime of anomalous diffusion. However from figure 3b one can clearly see that, with the fixed scale analysis, only two regimes survive: exponential and saturation. Comparing figure 3a and figure 3b one understands that the appearance of the spurious anomalous diffusion regime in the fixed time analysis is due to the mechanism described in section \ref{sec:2}. The absence of the diffusive regime $\lambda(\delta) \sim \delta^{-2}$ is due to the fact that the characteristic length of the velocity field, which is comparable with the typical distance between two close vortices, is not much smaller than the size of the basin. \subsection{Random walk on a fractal object: an anomalous diffusive case} \label{sec:3.3} In this section we discuss the case of particles performing a continuous random walk on a fractal object of fractal dimension $D_F$, where one has sub-diffusion. We show that also in situation of anomalous diffusion (e.g. sub-diffusion) the FSLE is able to recognize the correct behavior. In the following section we consider the case of fully developed turbulence which displays super-diffusion. In a fractal object due to the presence of voids, i.e. forbidden regions for the particles, one expects a decreasing of the spreading, and because of the self-similar structure of the domain (i.e. voids on all scales) a sub-diffusive behavior is expected. It is worth to note that the particles do not diffuse with the same law from any points of the domain (due to the presence of voids), hence in order to define a diffusive-like behavior one has to average over all possible particles' position. For discrete random walk on a fractal lattice it is known that the diffusion follows the law $<R^2(t)> \sim t^{2/D_W}$ with $D_W>2$, i.e. sub-diffusion \cite{rammal83}. The quantity $D_W$ is related to the spectral or {\it fracton} dimension $D_S$, by the relation $D_W=2 D_F/D_S$, and it depends on the detailed structure of the fractal object \cite{rammal83}. We study the relative dispersion of 2-D continuous random walk in a Sierpinsky Carpet with fractal dimension $D_F=\log\,8/\log\,3$. In our computation we use a resolution $3^{-5}$, i.e. the fractal is approximated by five steps of the recursive building rule, in practice we perform a continuous random walk in a basin obtained with the above approximation of the Sierpinsky Carpet. We initialize the particles inside one of the smallest resolved structures, then we follow the growth of the relative dispersion with the FSLE method, and redeploy the particles in a small cell randomly chosen at the beginning of each doubling time experiment. From fig.~4 one can see that $\lambda(\delta)\sim \delta^{-1/.45}$ which is an indication of sub-diffusion the exponent is in good agreement with the usual relative dispersion analysis (see the inset of fig.~4). \section{Application of the FSLE} \label{sec:4} \subsection{Drifter in the Adriatic Sea: data analysis and modelization} \label{sec:4.1} Lagrangian data recorded within oceanographic programs in the Mediterranean Sea \cite{Poulain98} offer the opportunity to apply the fixed scale analysis to a geophysical problem, for which the standard characterization of the dispersion properties gives poor information. The Adriatic Sea is a semi-enclosed basin, about 800 by 200 $km$ wide, connected to the whole Mediterranean Sea through the Otranto Strait \cite{Poulain98,ABPPRR97}. We adopt the reference frame in which the $x,y-$axes are aligned, respectively, with the short side (transverse direction), orthogonal to the coasts, and the long side (longitudinal direction), along the coasts. We have computed relative dispersion along the two axes, $\langle R^2_x(t)\rangle$, $\langle R^2_y(t) \rangle$ and FSLE $\lambda(\delta)$. The number of selected drifters for the analysis is 37, distributed in 5 different deployments in the Strait of Otranto, happened during the period December 1994 - March 1997, containing respectively 4, 9, 7, 7 and 10 drifters. To get as high statistics as possible, even to the cost of losing information on the seasonal variability, we shift the time tracks of all of the 37 drifters to $t-t_0$, where $t_0$ is the time of the deployment, so that the drifters can be treated as a whole cluster. Moreover, to restrict the analysis only to the Adriatic basin, we discarded a drifter as soon as its latitude goes south of $39.5$ N or its longitude goes beyond $19.5$ E. Before presenting the results of the data analysis, let us introduce a simplified model for the Lagrangian tracers motion in the Adriatic Sea. We assume as main features of the surface circulation the following elements \cite{Lacorataetal98}: the drifter motion is basically two-dimensional; the domain is a quasi-closed basin; an anti-clockwise coastal current; two large cyclonic gyres; some natural irregularities in the Lagrangian motion induced by small scale structures. On the basis of these considerations, we introduce a deterministic chaotic model with mixing properties for the Lagrangian drifters. The stream function is given by the sum of three terms: \begin{equation} \Psi(x,y,t)=\Psi_0(x,y) + \Psi_1(x,y,t) + \Psi_2(x,y,t) \label{4.1} \eeq defined as follows: \begin{equation} \Psi_0(x,y)={C_0 \over k_0} \cdot [- \sin(k_0 (y + \pi)) + \cos(k_0 (x + 2 \pi))] \label{4.2} \eeq \begin{equation} \Psi_i(x,y,t)={C_i \over k_i } \cdot \sin(k_i (x + \epsilon_i \sin(\omega_i t))) \sin(k_i (y + \epsilon_i \sin(\omega_i t + \phi_i))) \,,\hspace{10pt} (i=1,2)\,, \label{4.3} \eeq where $k_i \,=\, 2 \pi / \lambda_i$, for $i=0,1,2$, $\lambda_i$'s are the wavelengths of the spatial structures of the flow; analogously $\omega_j \,=\, 2 \pi / T_j$, for $j=1,2$, and $T_j$'s are the periods of the perturbations. In the non-dimensional expression of the equations, the length and time units have been set to $200 \; km$ and $7.5 \; days$, respectively. The stationary term $\Psi_0$ defines the boundary large scale circulation with positive vorticity. The contribution of $\Psi_1$ contains the two cyclonic gyres and it is explicitly time-dependent through a periodic perturbation. The term $\Psi_2$ gives the motion over scales smaller than the size of the large gyres and it is time-dependent as well. The zero-value isoline is defined as the boundary of the basin. According to observation, we have chosen the parameters so that the velocity range is around $\sim \; 0.3 \; m \, s^{-1}$; the length scales of the Eulerian structures are $L_B\sim \; 1000 \; km$ (coastal current), $L_0\sim \; 200 \; km$ (gyres) and $l_u\sim \; 50 \; km$ (vortices); the typical recirculation times, for gyres and vortices, are $\sim$ 1 month and $\sim$ 1 week, respectively; the oscillation periods are $ \simeq 10 \, days $ (gyres) and $ \simeq 2 \, days$ (vortices). Let us discuss now the comparison between data and model results. The relative dispersion along the two directions of the basin, for data and model trajectories, are shown in figures 5a,b. The results for the model are obtained from the spreading of a cluster of $10^4$ initial conditions. When a particle reaches the boundary ($\Psi = 0$) it is eliminated. For the diffusion properties, one cannot expect a scaling for $\langle R^{2}_{x,y}(t) \rangle$ before the saturation regime, since the Eulerian characteristic lengths are not too small compared with the basin size. Indeed, we do not observe a power law behavior neither for the experimental data nor for the numerical model. Let us stress that by opportunely fitting the parameters, we could obtain the model curves even closer to the experimental ones, but this would not be very meaningful since there is no clear theoretical expectation in a transient regime. Let us now discuss the finite size Lyapunov exponent. The analysis of the experimental data has been averaged over the total number of couples out of 37 trajectories, under the condition that the evolution of the distance between two drifters is no longer followed when any of the two exits the Adriatic basin. In fig. 6 we show the FSLE for data and model. In our case, as discussed above, we are far from asymptotical conditions, therefore we do not observe the scaling $\lambda(\delta) \sim \delta^{-2}$. The $\lambda_{M}(\delta)$ obtained from the minimal chaotic model (\ref{4.1}-\ref{4.3}) shows the typical step-like shape of a system with two characteristic time scales, and offers a scenario about how the FSLE of real trajectories may come out. The relevant fact is that the large-scale Lagrangian features are well reproduced, at least at a qualitative level, by a relatively simple model. We believe that this agreement is not due to a particular choice of the model parameters, but rather to the fact that transport is mainly dominated by large scales whereas small scale details play a marginal role. It is evident the major advantages of FSLE with respect to the usual fixed time statistics of relative dispersion: from the relative dispersion analysis of fig.~5 we are unable to recognize the underlying Eulerian structures, while the FSLE of fig.~6 suggests the presence of structures on different scales and with different characteristic times. In conclusion, the fixed scale analysis gives information for discriminating among different models for the Adriatic Sea. \subsection{Relative dispersion in fully developed turbulence} \label{sec:4.2} We consider now the relative dispersion of particles pairs advected by an incompressible, homogeneous, isotropic, fully developed turbulent field. The Eulerian statistics of velocity differences is characterized by the Kolmogorov scaling $\delta v(r) \sim r^{1/3}$, in an interval of scales $\l_u \ll r \ll L_0$, called the inertial range, $l_u$ is now the Kolmogorov scale. Due to the incompressibility of the velocity field particles will typically diffuse away from each other \cite{Cocke69,Orszag70}. For pair separations less than $l_u$ we have exponential growth of the separation of trajectories, typical of smooth flows, whereas at separations larger than $L_0$ normal diffusion takes place. In the inertial range the average pair separation is not affected neither by large scale components of the flow, which simply sweep the pair, nor by small scale ones, whose intensity is low and which act incoherently. Accordingly, the separation $R(t)$ feels mainly the action of velocity differences $\delta v(R(t))$ at scale $R$. As a consequence of the Kolmogorov scaling the separation grows with the {\em Richardson law} \cite{Richardson26,MY75} \begin{equation} \langle R^2(t) \rangle \sim t^{3} \, . \label{3.2.1} \eeq Non-asymptotic behavior takes place in such systems whenever $l_u$ is not much smaller than $L_0$, that is when the Reynolds number is not high enough. As a matter of facts even at very high Reynolds numbers, the inertial range is still insufficient to observe the scaling (\ref{3.2.1}) without any ambiguity. On the other hand, we shall show that FSLE statistics is effective already a relatively small Reynolds numbers. In order to investigate the problem of relative dispersion at various scale separations a practical tool is the use of synthetic turbulent fields. In fact, by means of stochastic processes it is possible to build a velocity field which reproduces the statistical properties of velocity differences observed in fully developed turbulence \cite{BBCCV98}. In order to avoid the difficulties related to the presence of sweeping in the velocity field, we limit ourselves to a correct representation of two-point velocity differences. In this case, if one adopts the reference frame in which one of the two tracers is at rest at the origin (the so called a Quasi-Lagrangian frame of reference), the motion of the second particle is ruled by the velocity difference in this frame of reference, which has the the same single time statistics of the Eulerian velocity differences \cite{LPP97,BCCV99}. The detailed construction of the synthetic Quasi-Lagrangian velocity field is presented in Appendix \ref{app:3}. In figure 7 we show the results of simulations of pair dispersion by the synthetic turbulent field with Kolmogorov scaling of velocity differences at Reynolds number $Re \simeq 10^{6}$ \cite{BCCV99}. The expected super-diffusive regime (\ref{3.2.1}) can be well observed only for huge Reynolds numbers (see also Ref. \cite{Maida}. To explain the depletion of scaling range for the relative dispersion let us consider a series of pair dispersion experiments, in which a couple of particles is released at a separation $R_0$ at time $t=0$. At a fixed time $t$, as customarily is done, we perform an average over all different experiments to compute $\langle R^2(t) \rangle$. But, unless $t$ is large enough that all particle pairs have ``forgotten'' their initial conditions, the average will be biased. This is at the origin of the flattening of $\langle R^2(t) \rangle$ for small times, which we can call a crossover from initial condition to self similarity. In an analogous fashion there is a crossover for large times, of the order of the integral time-scale, since some couples might have reached a separation larger than the integral scale, and thus diffuse normally, meanwhile other pairs still lie within the inertial range, biasing the average and, again, flattening the curve $\langle R^2(t) \rangle$. This correction to a pure power law is far from being negligible for instance in experimental data where the inertial range is generally limited due to the Reynolds number and the experimental apparatus. For example, references \cite{FV98,Fung92} show quite clearly the difficulties that may arise in numerical simulations with the standard approach. To overcome these difficulties we exploit the approach based on the fixed scale statistics. The outstanding advantage of averaging at a fixed separation scale is that it removes all crossover effects, since all sampled pairs belong to the inertial range. The expected scaling properties of the doubling times is obtained by a simple dimensional argument. The time it takes for particle separation to grow from $R$ to $2 R$ can be estimate as $T(R) \sim R/\delta v(R)$; we thus expect for the inverse doubling times the scaling \begin{equation} \left\langle {1 \over T(R)} \right\rangle \sim {\langle \delta v(R) \rangle \over R} \sim R^{-2/3} \label{eq:5.1} \end{equation} In figure 8 the great enhancement of the scaling range achieved by using the doubling times is evident. In addition, by using the FSLE it is possible to study in details the effect of Eulerian intermittency on the Lagrangian statistics of relative dispersion. See Ref.\cite{BCCV99} for a detailed discussion and a comparison with a multifractal scenario. The conclusion that can be drawn is that in this case doubling time statistics makes it possible a much better estimate of the scaling exponent with respect to the standard -- fixed time -- statistics. \section{Conclusions} \label{sec:5} In the study of relative dispersion of Lagrangian tracers one has to tackle situations in which the asymptotic behavior is never attained. This may happen in presence of many characteristic Eulerian scales or, what is typical of real systems, in presence of boundaries. It is worth to stress that such kind of systems are very common in geophysical flows \cite{D}, and also in plasma physics \cite{sabot}. Therefore a close understanding of non-asymptotic transport properties can give much relevant information about these natural phenomena. To face these problems, in recent years, there have been proposed different approaches whose common ingredient is basically an ``exit time'' analysis. We remind the symbolic dynamics \cite{A,B} and the chaotic scattering \cite{C} approaches, the exit time description for transport in semi-enclosed basins \cite{D}, symplectic maps \cite{meiss}, open flows \cite{E} and in plasma physics \cite{sabot}. In this paper we have discussed the applications of the Finite Size Lyapunov Exponent, $\lambda(\delta)$, in the analysis of several situations. This method is based on the identification of the typical time $\tau(\delta)$ characterizing the diffusive process at scale $\delta$ through the exit time. This approach is complementary to the traditional one, in which one looks at the average size of the clouds of tracers as function of time. For values of $\delta$ much smaller than the smallest characteristic length of the Eulerian velocity field, one has that $\lambda(\delta)$ coincides with the maximum Lagrangian Lyapunov Exponent. For larger $\delta$ the shape of $\lambda(\delta)$ depends on the detailed mechanisms of spreading, i.e. the structure of the advecting velocity field and/or the presence of boundaries. The diffusive regime corresponds to the behavior $\lambda(\delta)\simeq D/\delta^2$. If $\delta$ gets close to its saturation value, i.e. the characteristic size of the basin, the universal shape of $\lambda(\delta)$ can be obtained on the basis of dynamical system theory. In addition, we have shown that the fixed scale method is able to recognize the presence of a genuine anomalous diffusion. A remarkable advantage of working at fixed scale (instead of at fixed time as in the traditional approach) is its ability to avoid misleading results, for instance apparent anomalous scaling over a certain time interval. Moreover, with the FSLE one obtains the proper scaling laws also for a relatively small inertial range for which the standard technique gives rather controversial answers. The proposed method can be also applied in the analysis of drifter experimental data or in numerical model for Lagrangian transport. \section{Acknowledgments} We thank V. Artale, E. Aurell, L. Biferale, P. Castiglione, A. Crisanti, M. Falcioni, R. Pasmanter, P.M. Poulain, M. Vergassola and E. Zambianchi for collaborations and discussions in last years. A particular acknowledgment to B. Marani for the continuous and warm encouragement. We are grateful to the ESF-TAO ({\it Transport Processes in the Atmosphere and the Oceans}) Scientific Program for providing meeting opportunities. This paper has been partly supported by INFM (Progetto di Ricerca Avanzato PRA-TURBO), MURST (no. 9702265437), and the European Network {\it Intermittency in Turbulent Systems} (contract number FMRX-CT98-0175).
\section{Introduction} The theory of standard bases has developed since B.~Buchberger introduced standard bases of ideals in polynomial rings with respect to semigroup wellorderings (now called Gr\"obner bases) in 1965 \cite{buchberger}. It has been extended to certain localisations of the polynomial rings by allowing semigroup orderings which are not wellorderings \cite{mora,graebe,greuelpfister,moraseven}. The main tool to compute such a standard basis is the Buchberger algorithm \cite{buchberger,beckerweispfenning}. By choosing an appropriate semigroup ordering, invariants of ideals may be computed from a standard basis with respect to this ordering. For example, the Hilbert function can be obtained from a standard basis with respect to a degree ordering. However there are some invariants which are not related to semigroup orderings. Consider an isolated hypersurface singularity $f\colon\mathbb{C}^n\rightarrow\mathbb{C}$ with nondegenerate principal part. Its Milnor number $\mu(f)$ is readily computed from a standard basis of the Jacobian ideal $J_f$ of $f$ with respect to any local semigroup ordering (in fact with respect to any local ordering as will be shown). However the spectrum of $f$ is the Poincare series of ${\cal O}_{\mathbb{C}^n,\mathbf{0}}/J_f$ graded with respect to the Newton filtration given by $f$ \cite{saito,kv}. The Newton filtration can be refined to a local semigroup ordering if and only if $f$ is semiquasihomogeneous. Here we study standard bases with respect to orderings which are refinements of Newton filtrations. In general, such an ordering is not a semigroup ordering. In the first section normal and noetherian orderings are introduced. This is the class of orderings for which we are able to give a normal form algorithm which terminates (algorithm \ref{algorithm:normalform}). The setup for standard bases with respect to normal orderings is outlined in the second section. To compute standard bases, $s$-polynomial sets and reducing sets are introduced. With their help a modified Buchberger algorithm can be formulated (algorithm \ref{algorithm:standardbasis}). It terminates for any normal noetherian ordering for which reducing sets exist (theorem \ref{theorem:standardbasis}) and returns a standard basis (theorem \ref{theorem:standardbasisx}). In the next section Newton orderings are introduced as refinements of Newton filtrations. Some Newton orderings are not normal (lemma \ref{lemma:newtonnormal}), but they all are noetherian and admit reducing sets (propositions \ref{proposition:newtonnoether} and \ref{proposition:newtonred}). The case of a zero dimensional ideal and a local ordering is studied in in the fourth section. Here standard bases do always exist (proposition \ref{proposition:always}). Moreover a standard basis with respect to a given local ordering can be computed from a standard basis with respect to another local ordering using only linear algebra (algorithm \ref{algorithm:convert}). In the last section we show how to compute the spectrum of an isolated complex hypersurface singularity with nondegenerate principal part (corollary \ref{corollary:spectrum}). The author implemented the computation of the spectrum into the computer algebra program {\tt Singular} \cite{singular} using algorithm \ref{algorithm:convert}. \section{Orderings}\label{sect:orderings} First fix some notation. Let $\mathbb{N}=\{0,1,2,\ldots\}$ be the nonnegative integers, $K$ a field and $K[\bx]=K[x_1,\ldots,x_n]$ the polynomial ring in $n\geq 1$ indeterminates over $K$. We use the exponent notation $\bx^\balpha= x_1^{\alpha_1}\cdots x_n^{\alpha_n}$, ${\pmb{\alpha}}=(\alpha_1,\ldots,\alpha_n)\in\mathbb{N}^n$. Let \begin{align*} % \mathscr{M} &= \{\bx^\balpha\mid{\pmb{\alpha}}\in\mathbb{N}^n\}\text{\ be the set of all monomials of $K[\bx]$ and}\\ \mathscr{T} &= \{c\,\bx^\balpha|c\in K^\ast, \bx^\balpha\in\mathscr{M}\}\text{\ the set of all terms in $K[\bx]$.} % \end{align*} The polynomials of $K[\bx]$ are sums of the form $\sum_{{\pmb{\alpha}}\in A}c_{\pmb{\alpha}}\bx^\balpha$, where $A\subset\mathbb{N}^n$ is a finite set and $c_{\pmb{\alpha}}\bx^\balpha\in\mathscr{T}$ for all ${\pmb{\alpha}}\in A$ (we do not write monomials with zero coefficient). Consider a total ordering $\prec$ on $\mathbb{N}^n$ (which is not not necessarily a semigroup ordering) and denote the induced ordering on $\mathscr{M}$ also by $\prec$. \begin{definition} % Let $G={\{\fele}\}\subsetK[\bx]$ be a finite set of polynomials, $I\subseteqK[\bx]$ an ideal and $f=\sum_{{\pmb{\alpha}}\in A}c_{\pmb{\alpha}}\bx^\balpha\in K[\bx]$. % \begin{itemize} % \item[1)] $LM(f)=\mathbf{x}^{\max A}$ is called the {\em lead monomial of $f$}. \item[2)] $LC(f)=c_{\max A}$ is called the {\em lead coefficient of $f$}. \item[3)] $LT(f)=LC(f)\cdot LM(f)$ is called the {\em lead term of $f$}. \item[4)] $L(G)=\{LM(\bx^\balpha g)\mid \bx^\balpha\in\mathscr{M}, g\in G\}$ is called the {\em lead monomial set of $G$}. \item[5)] $L(I)=\{LM(g)\mid g\in I\}$ is called the {\em lead monomial set of $I$}. % \end{itemize} % \end{definition} Any nonempty set of monomials which is the lead monomial set of a finite set of polynomials or of an ideal will be called {\em lead monomial set}. For a semigroup ordering $L(I)$ and $L(G)$ can be identified with monomial ideals. In general this is not the case. But the ordering $\prec$ still satisfies \begin{equation*} % LM(f+g)\preccurlyeq\max\left( LM(f),LM(g)\right) % \end{equation*} for all $f$, $g\in K[\bx]$. Instead of respecting the semigroup structure of $\mathbb{N}^n$, the ordering $\prec$ should satisfy the following condition. \begin{definition} % The ordering $\prec$ is called {\em normal}, if for all $f\in K[\bx]$, $\bx^\balpha\in\mathscr{M}$ % \begin{align*} % \bx^\balpha\prec 1 \Longrightarrow LM(\bx^\balpha f) \prec LM(f). % \end{align*} % A normal ordering $\prec$ is called {\em global} (resp.~{\em local}) if $x_i\succ 1$ (resp.~$x_i\prec 1$), $1\leq i\leq n$. A normal ordering is called {\em mixed} if it is neither global nor local. % \end{definition} For a normal ordering the set $S_{\mlt}=\{g\inK[\bx]\mid LM(g)=1\}$ is multiplicatively closed. Moreover $LM(f)=LM(gf)$ for all $f$, $g\inK[\bx]$ with $LM(g)=1$. For semigroup orderings the lead monomial sets behave like ideals (in a noetherian ring), i.e.~every increasing sequence gets stationary. \begin{definition} % The ordering $\prec$ is called {\em noetherian} if the following condition is satisfied: % \begin{itemize} % \item[] Every increasing sequence of lead monomial sets $L_0\subseteq L_1\subseteq L_2\subseteq\ldots$ gets stationary. % \end{itemize} % \end{definition} Any semigroup ordering is normal and noetherian, but the converse does not hold. \section{Standard bases}\label{sect:general} Proceeding along the lines of \cite{greuelpfister}, we introduce standard bases with respect to normal orderings. From now on let $\prec$ be a normal ordering. Let $S_\loc^{-1}\Kx$ be the localisation of $K[\bx]$ in $S_{\mlt}$. For every ideal $I\subseteqK[\bx]$, let $S_\loc^{-1}I = I\otimes_{K[\bx]}S_\loc^{-1}\Kx$ and $\overline{I}=S_\loc^{-1}I\capK[\bx]$. Note that always $L(I)=L(\overline{I})$. \begin{definition}\label{definition:standard} % Let $I\subseteqK[\bx]$ be an ideal and $G={\{\fele}\}\subseteq\overline{I}$. % \begin{itemize} % \item[1)] $G$ is called a {\em standard basis of $I$} if $L(G)=L(I)$. \item[2)] $G$ is called {\em interreduced} if $L(\{f_i\})\not\subseteq L(G\setminus\{f_i\})$ for all $i=1,\ldots,r$. \item[3)] $G$ is called {\em reduced} if for every $f\in G$ no monomial of $f$ except its lead monomial is contained in $L(G)$. % \end{itemize} % \end{definition} If $\prec$ is not a semigroup ordering, then in general not every ideal has a standard basis. Albeit for noetherian orderings this is true. \begin{lemma} % If $\prec$ is noetherian, the every ideal $I\subseteqK[\bx]$ has a standard basis. % \end{lemma} {\noindent\it Proof:\ } The standard proof applies. $\square$\medskip\par Every standard basis $G$ of $I$ can be shortened to an interreduced standard basis by iteratively deleting those $f\in G$ for which $L(\{f\})\subseteq L(G\setminus\{f\})$. Reduced standard bases do in general only exist for global orderings. To do standard basis computations a normal form is needed: \begin{definition}\label{definition:normal_form} % \cite[def.~1.5]{greuelpfister} Let $\mathscr{F}=\{G\subset K[\bx]\mid G\text{\ ordered and finite}\}$. A function $NF\colon K[\bx]\times\mathscr{F}\rightarrow K[\bx]$ is called {\em normal form} if for all $p\in K[\bx]$ and $G\in\mathscr{F}$ $NF(p,G)\neq 0 \Rightarrow LM(NF(p,G))\not \in L(G)$. Then $NF(p,G)$ is called {\em normal form of $p$ with respect to $G$}. % \end{definition} Following \cite{greuelpfister}, to get a normal form which does not only work in the global case, one considers homogeneous polynomials in $K[t,\mathbf{x}]$. So let \begin{align*} % \mathscr{M}_t &=\{t^{\alpha_0}\bx^\balpha\mid \alpha_0\in\mathbb{N}, {\pmb{\alpha}}\in\mathbb{N}^n\},\\ \mathscr{T}_t &=\{c\,t^{\alpha_0}\bx^\balpha\mid c\in K^\ast, \alpha_0\in\mathbb{N}, {\pmb{\alpha}}\in\mathbb{N}^n\}, % \end{align*} be the monomials and terms of $K[t,\bx]$. Define a global wellordering $\prec_h$ on $\mathscr{M}_t$ by setting \begin{align*} % t^{\alpha_0}\bx^\balpha\prec_h t^{\beta_0}\bx^\bbeta\Longleftrightarrow & \ \alpha_0 +|{\pmb{\alpha}}| < \beta_0+|{\pmb{\beta}}|\quad\text{or}\\ &\ \alpha_0 +|{\pmb{\alpha}}| = \beta_0+|{\pmb{\beta}}|\quad\text{and}\quad\bx^\balpha\prec\bx^\bbeta. % \end{align*} With respect to powers of $t$, this ordering behaves like a semigroup ordering. We frequently need to consider lead monomials in $K[t,\bx]$ and $K[\bx]$. For every polynomial $f=\sum_{{\pmb{\alpha}}\in A}c_{\pmb{\alpha}}\bx^\balpha\in K[\bx]$, let $f^h=\sum_{{\pmb{\alpha}}\in A}c_{\pmb{\alpha}} t^{\deg f-|{\pmb{\alpha}}|}\bx^\balpha$ be the homogenisation with respect to $t$. Conversely, for $f=\sum_{(\alpha_0,{\pmb{\alpha}})\in A} c_{(\alpha_0,{\pmb{\alpha}})}t^{\alpha_0}\bx^\balpha$ let $\left.f\right|_{t=1}= \sum_{(\alpha_0,{\pmb{\alpha}})\in A} c_{(\alpha_0,{\pmb{\alpha}})}\bx^\balpha$ be its dehomogenisation. For polynomials of $K[\bx]$ (resp.~$K[t,\bx]$) we always use $\prec$ (resp.~$\prec_h$) in the computation of lead monomials, lead terms, \ldots . Let $G={\{\fele}\}\subsetK[t,\bx]$ be a finite ordered set of homogeneous polynomials and let $p\inK[t,\bx]$ homogeneous. For a semigroup ordering, the following is just the normal form {\bf NFMora} of \cite{greuelpfister}. \begin{algorithm}\label{algorithm:normalform} % $h:={\rm\bf NormalForm}(p,G)$\\ \hspace*{1cm} $h:=p$\\ \hspace*{1cm} $H:=\emptyset$\\ \hspace*{1cm} WHILE $\left(\begin{array}{l} \text{exist $f\in G\cup H$, $\eta\in\mathscr{T}_t$, $\alpha\in\mathbb{N}$ with}\\ \text{$LT(\eta f)=LT(t^\alpha h)$ and $\eta\preccurlyeq 1$ if $f\in H$} \end{array}\right)$ DO\\ \sa\sa choose first such $f$ with $\alpha$ minimal\\ \sa\sa IF $\alpha>0$ THEN\\ \sa\sa\sa $H:=H\cup\{h\}$\\ \sa\sa $h:=t^\alpha h-\eta f$\\ \sa\sa IF $t\mid h$ THEN\\ \sa\sa\sa choose $\alpha$ maximal with $t^\alpha\mid h$\\ \sa\sa\sa $h:=h/t^\alpha$ % \end{algorithm} \begin{lemma}\label{lemma:normal_form} % If $h$ is a normal form of $p$ with respect to $G={\{\fele}\}$ computed by {\rm\bf NormalForm}, then there exist $\xi_1,\ldots,\xi_k\in\mathscr{T}_t$, $g\inK[t,\bx]$ homogeneous and $j_1,\ldots,j_k\in\{1,\ldots,r\}$ such that % \begin{equation*} % gp=\sum_{i=1}^k\xi_if_{j_i} + h % \end{equation*} % with $LM(g|_{t=1})=1$ and $LM(h|_{t=1})\not\in L(G|_{t=1})$. Moreover $gp$, $\xi_1 f_{j_1},\ldots,\xi_k f_{j_k}$ and $h$ (if nonzero) are homogeneous of the same degree with % \begin{equation*} % LM(gp) = LM(\xi_1f_{j_1})\succ_h\ldots\succ_h LM(\xi_k f_{j_k})\succ_h LM(h). % \end{equation*} % \end{lemma} {\noindent\it Proof:\ } This just the proof of \cite[theorem 1.9 2)]{greuelpfister}. $\square$\medskip\par \begin{proposition}\label{proposition:normal_form} % If the ordering $\prec$ is noetherian, then {\rm\bf NormalForm}\ is a normal form in the sense of definition \ref{definition:normal_form}. % \end{proposition} {\noindent\it Proof:\ } In view of lemma \ref{lemma:normal_form} it is enough to show that {\rm\bf NormalForm}\ terminates. But $\prec$ is noetherian, so the proof of \cite[prop.~1.9 1)]{greuelpfister} applies. $\square$\medskip\par To actually compute such a standard basis one needs $s$-polynomials. \begin{definition} % Let $f=\sum_{{\pmb{\alpha}}\in A}c_{\pmb{\alpha}}\bx^\balpha$, $g=\sum_{{\pmb{\beta}}\in B}d_{\pmb{\beta}}\bx^\bbeta\in K[\bx]$. Then % \begin{equation*} % \operatorname{spoly} (f,g)_{({\pmb{\alpha}},{\pmb{\beta}})}= \left(d_{\pmb{\beta}}\bx^\bbeta f-c_{\pmb{\alpha}}\bx^\balpha g\right)/\gcd\left(\bx^\balpha,\bx^\bbeta\right) % \end{equation*} % is called {\em $s$-polynomial of $f$ and $g$ at $({\pmb{\alpha}},{\pmb{\beta}})$}. Moreover % \begin{align*} % \operatorname{Spoly}(f,g)=\left\{ \operatorname{spoly}(f,g)_{({\pmb{\alpha}},{\pmb{\beta}})}\mid \begin{array}{l} ({\pmb{\alpha}},{\pmb{\beta}})\in A\times B, \exists \mathbf{x}^{\pmb{\gamma}}\in\mathscr{M}:\\ LM(\mathbf{x}^{\pmb{\gamma}}\operatorname{spoly}(f,g)_{({\pmb{\alpha}},{\pmb{\beta}})})= \mathbf{x}^{\pmb{\gamma}}\operatorname{lcm}(\bx^\balpha,\bx^\bbeta) \end{array} \right\} % \end{align*} % is called the {\em $s$-polynomial set of $f$ and $g$}. % \end{definition} For a semigroup ordering $\left|\operatorname{Spoly}(f,g)\right|=1$ and we get the usual $s$-polynomial. Now we want to imitate the Buchberger algorithm. Given a finite set of generators $G={\{\fele}\}$ of an ideal $I$, this algorithm enlarges $G$ by some elements of $I$ such that all $s$-polynomials $\operatorname{spoly}(f,g)$, $f,g\in G$ reduce to zero. For a semigroup ordering this implies that $\bx^\balpha\operatorname{spoly}(f,g)$ reduces to zero for all $\bx^\balpha\in\mathscr{M}$. In our setup this is not the case. Let $G={\{\fele}\}\subsetK[t,\bx]$ be a finite set of homogeneous polynomials and let $f\in I=(f_1,\ldots,f_r)$ homogeneous. Assume that that {\rm\bf NormalForm}\ terminates for $\prec$. \begin{definition}\label{definition:reducing_set} % A finite set $R(f,G)\subset\mathscr{M}$ of monomials is called a {\em reducing set for $(f,G)$} if the following holds: For all $\bx^\balpha\in\mathscr{M}$, there exist % \begin{itemize} % \item a homogeneous polynomial $g\inK[t,\bx]$ with $LM(g|_{t=1})=1$, \item $\xi_1,\ldots,\xi_k\in\mathscr{T}_t$ and \item $h_1,\ldots,h_k\in G\cup\{{\rm\bf NormalForm}(\bx^\bbeta f,G)\mid \bx^\bbeta\in R(f,G)\}$ % \end{itemize} % such that the polynomial $\bx^\balpha f$ has a representation of the form % \begin{equation*} % g\bx^\balpha f = \sum_{i=1}^k\xi_i h_i % \end{equation*} % with $LM(g\bx^\balpha f)=LM(\xi_1 h_1)\succ_h LM(\xi_i h_i)$, $2\leq i\leq k$. If there exists a reducing set for every $(f,G)$ as above, then we say that {\em reducing sets exist}. % \end{definition} For a semigroup ordering we always can take $R(f,G)=\{1\}$. Now we can formulate a standard basis algorithm. Let $G={\{\fele}\}\subsetK[\bx]$ be a finite set of polynomials. Let $\prec$ be such that {\rm\bf NormalForm}\ terminates and that reducing sets exist. Moreover assume that the procedure ${\rm\bf ReducingSet}(h,S)$ computes a reducing set for $(h,S)$. \par \begin{algorithm}\label{algorithm:standardbasis}\ \newline\noindent % $S:={\rm\bf StandardBasis}(G)$\\ \hspace*{1cm}$S:=G^h$\\ \hspace*{1cm}$P:=\{(f,g)\mid f,g\in S\}$\\ \hspace*{1cm}WHILE $P\neq\emptyset$ DO\\ \hspace*{2cm}choose $(f,g)\in P$; $P:=P\setminus\{(f,g)\}$\\ \hspace*{2cm}FOR ALL $h\in\operatorname{Spoly}(f,g)$ DO\\ \hspace*{3cm}R:={\rm\bf ReducingSet}(h,S)\\ \hspace*{3cm}FOR ALL $\bx^\balpha\in R$ DO\\ \hspace*{4cm}$p:={\rm\bf NormalForm}(\bx^\balpha h,S)$\\ \hspace*{4cm}IF $p\neq 0$ THEN\\ \hspace*{5cm}$S:=S\cup\{p\}$\\ \hspace*{5cm}$P:=P\cup\{(p,f)\mid f\in S\}$\\ \hspace*{1cm}$S:=S|_{t=1}$ % \end{algorithm} \begin{theorem}\label{theorem:standardbasis} % Let $\prec$ be noetherian such that reducing sets exist. Then {\rm\bf StandardBasis}\ terminates. % \end{theorem} {\noindent\it Proof:\ } Again the standard proof applies. $\square$\medskip\par \begin{theorem}\label{theorem:standardbasisx} % Let $\prec$ be such that {\rm\bf NormalForm}\ terminates and reducing sets exist. Let $I\subseteqK[\bx]$ be an ideal. Equivalent for $G={\{\fele}\}\subseteq\overline{I}$ are: % \begin{itemize} % \item[ i)] $G$ is a standard basis of $I$. \item[ ii)] $G={\rm\bf StandardBasis}(G)$. \item[iii)] ${\rm\bf NormalForm}(\bx^\balpha g^h,G^h)=0$ for all $\bx^\balpha\in\mathscr{M}$, $h\in\operatorname{Spoly}(f_i,f_j)$, $1\leq i\leq j\leq r$. \item[ vi)] $f\in\overline{I}\Leftrightarrow{\rm\bf NormalForm}(f^h,G^h)=0$. \item[ v)] $f\in\overline{I}\Leftrightarrow gf=\sum_{i=1}^k\xi_i f_{j_i}$ for some $g\inK[\bx]$, $\xi_1,\ldots,\xi_k\in\mathscr{T}$ and $j_1,\ldots,j_k\in\{1,\ldots,r\}$ with $LM(g)=1$ and $LM(f) = LM(\xi_1 f_{j_1}) \succ\ldots\succ LM(\xi_k f_{j_k})$. % \end{itemize} % \end{theorem} {\noindent\it Proof:\ } The standard proof applies to {\it i)}$\Rightarrow${\it ii)}, {\it iii)}$\Rightarrow${\it iv)}, {\it iv)}$\Rightarrow${\it v)} and {\it v)}$\Rightarrow${\it i)}. {\it ii)}$\Rightarrow${\it iii)}: Assume that {\it ii)} holds and let $f\inK[t,\bx]$ homogeneous. First we show the following\\ {\bf Claim:} If $f$ has a representation of the form \begin{equation*} % f = \sum_{i=1}^k\xi_i f_{j_i}\quad\text{with}\quad LM(f) = LM(\xi_1 f_{j_1}) \succ_h LM(\xi_i f_{j_i}),\quad 2\leq i\leq k, % \end{equation*} then $\tilde{f}=f-\xi_1 f_{j_1}$ has a representation of the form \begin{equation*} % g\tilde{f} = \sum_{i=1}^{\tilde{k}}\tilde{\xi}_i f_{\tilde{j}_i}\quad\text{with}\quad LM(g\tilde{f}) = LM(\tilde{\xi}_1 f_{\tilde{j}_1})\succ_h LM(\tilde{\xi}_i f_{\tilde{j}_i}),\quad 2\leq i\leq \tilde{k}, % \end{equation*} for a homogeneous $g\in K[t,\bx]$ with $LM(g|_{t=1})=1$. Indeed, after a permutation of summands $\tilde{f}=\sum_{i=2}^k\xi_i f_{j_i}$ with $LM(\tilde{f})=LM(\xi_2 f_{j_2})=\ldots=LM(\xi_l f_{j_l})\succ_h LM(\xi_i f_{j_i})$, $l+1\leq i\leq k$ for a $l\in\{1,\ldots,k\}$. If $l=1$ we are done. Otherwise let $c_i=LC(\xi_i f_{j_i})$, then \begin{align*} % \sum_{i=2}^l\xi_i f_{j_i} &= \underset{\text{lead terms cancel}}{\underbrace{\xi_2f_{j_2}-\frac{c_2}{c_3}\xi_3f_{j_3}}}+ \left(1+\frac{c_2}{c_3}\right)\xi_3f_{j_3}+ \sum_{i=4}^l\xi_i f_{j_i} \\ &= \eta h +\left(1+\frac{c_2}{c_3}\right)\xi_3 f_{j_3}+ \sum_{i=4}^l\xi_i f_{j_i} % \end{align*} for some $\eta\in\mathscr{T}_t$ and $h\in\operatorname{Spoly}(f_{j_2},f_{j_3})$. Since {\it ii)} holds $\eta h$ has a representation $g\eta h=\sum_{i'=1}^{k'}\xi_i'f_{j_i'}$ with $LM(g\xi_2f_{j_2})\succ_h LM(g\eta h)=LM(\xi_1'f_{j_1'})\succ_h LM(\xi_i'f_{j_i'})$, $2\leq i\leq k'$. Then \begin{align*} % g\tilde{f} &= \left(1+\frac{c_2}{c_3}\right)g\xi_3f_{j_3} + \sum_{i=4}^kg\xi_i f_{j_i}+ \sum_{i=1}^{k'}\xi_i'f_{j_i'} % \end{align*} and $LM(g\tilde{f})=LM(g\xi_3f_{j_3})=\ldots=LM(g\xi_l f_{j_l})$ and $LM(g\tilde{f})\succ_h LM(\xi_igf_{j_i})$, $l+1\leq i\leq k$ and $LM(g\tilde{f})\succ_h LM(g\xi_i'f_{j'_i})$, $1\leq i\leq k'$. Thus (after expansion) we have found a representation of $g\tilde{f}$ with $l-1$ lead terms instead of $l$. Then the claim follows by induction. Now let $\bx^\balpha\in\mathscr{M}$ and let $h\in\operatorname{Spoly}(f_i,f_j)$. Then there exists a $g\in K[t,\bx]$ with lead monomial 1 such that $g\bx^\balpha h$ has a representation as in the claim. Then the claim shows that during the computation of ${\rm\bf NormalForm}(\bx^\balpha h,G)$ on never ends up with an element which cannot be reduced except zero. Thus ${\rm\bf NormalForm}(\bx^\balpha h,G)=0$ and {\it iii)} follows. $\square$\medskip\par \begin{corollary} % Let $\prec$ be such that {\rm\bf NormalForm}\ terminates and reducing sets exist. If $G$ is a standard basis of the ideal $I\subseteqK[\bx]$, then $G$ generates $S_\loc^{-1}I$ over $S_\loc^{-1}\Kx$. % \end{corollary} \section{The Newton ordering}\label{sect:newton} Let $\mathscr{L}$ be a nonempty finite set of nonzero linear forms $l\colon\mathbb{R}^n\rightarrow\mathbb{R}$ and let ${\pmb{\delta}}\in(\mathbb{R}_0^+)^n$ be fixed. We call the real number \begin{align*} % w_{\pmb{\delta}}({\pmb{\alpha}}) &= \min\{l({\pmb{\alpha}}+{\pmb{\delta}})\mid l\in L\} % \end{align*} the {\em weight} of ${\pmb{\alpha}}\in(\mathbb{R}_0^+)^n$ with respect to $\mathscr{L}$. Let $C_{\pmb{\delta}}(l)=\{{\pmb{\alpha}}\in(\mathbb{R}_0^+)^n\mid w_{\pmb{\delta}}({\pmb{\alpha}})= l({\pmb{\alpha}}+{\pmb{\delta}})\}$, $l\in\mathscr{L}$. A subscript of ${\pmb{\delta}}=\mathbf{0}$ will be omitted. Then $C_{\pmb{\delta}}(l)$ is a cone with respect to $C(l)$: for all ${\pmb{\alpha}}\in C_{\pmb{\delta}}(l)$, ${\pmb{\beta}}\in C(l)$ by definition $w_{\pmb{\delta}}({\pmb{\alpha}}+{\pmb{\beta}})=l({\pmb{\alpha}}+{\pmb{\beta}})$, so ${\pmb{\alpha}}+{\pmb{\beta}}\in C_{\pmb{\delta}}(l)$. \begin{definition} % $\mathscr{L}$ is called {\em rational} if all cones $C(l)$, $l\in\mathscr{L}$ are rational (i.e.~every cone is the locus where a finite set of linear forms with rational coefficients is nonnegative). % \end{definition} For the rest of this section let $\mathscr{L}$ be rational and ${\pmb{\delta}}\in(\mathbb{Q}_0^+)^n$. Restricting $w_{\pmb{\delta}}(.)$ to $\mathbb{N}^n$ and identifying $\mathscr{M}$ with $\mathbb{N}^n$ we call \begin{equation*} % w_{\pmb{\delta}}(\bx^\balpha) = \min\{l({\pmb{\alpha}}+{\pmb{\delta}})\mid l\in\mathscr{L}\} % \end{equation*} the {\em Newton weight} of $\bx^\balpha\in\mathscr{M}$. Define the Newton weight of $f=\sum_{{\pmb{\alpha}}\in A}c_{\pmb{\alpha}}\bx^\balpha$ to be \begin{equation*} % w_{\pmb{\delta}}(f) =\min\{w_{\pmb{\delta}}(\bx^\balpha)\mid{\pmb{\alpha}}\in A\}. % \end{equation*} The induced filtration $K[\bx]_s=\{f\in K[\bx]\mid w_{\pmb{\delta}}(f)\geq s\}$ of $K[\bx]$ is called a {\em Newton filtration}. This filtration does not distinguish all monomials in general. So take any semigroup order $\prec_0$ and define \begin{align*} % \bx^\balpha\prec\bx^\bbeta\Longleftrightarrow & \ w_{\pmb{\delta}}(\bx^\balpha) > w_{\pmb{\delta}}(\bx^\bbeta)\quad\text{or}\\ &\ w_{\pmb{\delta}}(\bx^\balpha) = w_{\pmb{\delta}}(\bx^\bbeta)\quad\text{and}\quad \bx^\balpha\prec_0\bx^\bbeta. % \end{align*} for $\bx^\balpha$, $\bx^\bbeta\in\mathscr{M}$. This ordering is called a {\em Newton ordering}. A Newton filtration (resp.~a Newton ordering) is a filtration (resp.~an ordering) which arises in the above way. \begin{lemma}\label{lemma:newtonnormal} % Let $\prec$ be a Newton ordering. Then $\prec$ is normal in the following three cases: % \begin{itemize} % \item[1)] All linear forms $l\in\mathscr{L}$ have nonnegative coefficients. \item[2)] All linear forms $l\in\mathscr{L}$ have nonpositive coefficients. \item[3)] ${\pmb{\delta}}=0$. % \end{itemize} % \end{lemma} {\noindent\it Proof:\ } The first two cases are obvious. So let ${\pmb{\delta}}=\mathbf{0}$ and $\bx^\balpha$, $\bx^\bbeta\in\mathscr{M}$. We have to show that $\bx^\bbeta\prec 1$ implies $\bx^\balpha\bx^\bbeta\prec\bx^\balpha$. There exist linear forms $l$, $l'$, $l''\in\mathscr{L}$ such that $w(\bx^\balpha)=l({\pmb{\alpha}})$, $w(\bx^\bbeta)=l'({\pmb{\beta}})$ and $w(\bx^\balpha\bx^\bbeta)=l''({\pmb{\alpha}}+{\pmb{\beta}})$. Therefore \begin{align*} % w(\bx^\balpha\bx^\bbeta) &= l''({\pmb{\alpha}}+{\pmb{\beta}}) = l''({\pmb{\alpha}})+l''({\pmb{\beta}}) \\ &\geq l({\pmb{\alpha}})+l'({\pmb{\beta}}) = w(\bx^\balpha)+w(\bx^\bbeta). % \end{align*} Then $\bx^\bbeta\prec 1$ implies either $w(\bx^\bbeta)< 0\Rightarrow w(\bx^\balpha\bx^\bbeta)<w(\bx^\balpha)$ or $w(\bx^\bbeta)=0$ and ${\pmb{\beta}}\prec_0 1\Rightarrow w(\bx^\balpha\bx^\bbeta)=w(\bx^\balpha)$ and $\bx^\balpha\bx^\bbeta\prec_0\bx^\balpha$. In both cases we get $\bx^\bbeta\bx^\balpha\prec\bx^\balpha$. $\square$\medskip\par Note that $\prec$ is an ordering on $\mathscr{M}$ which is not a semigroup ordering in general. For example let $\mathscr{L}=\{x+2y,2x+y\}$ and $f=x+y$ (figure \ref{figure:newta}). \begin{figure}[H] % \centering \setlength{\unitlength}{1mm} \begin{picture}(55,55)(0,0) % \put(0,0){\epsfig{file=figure1.eps,height=5cm}} \put(52,0){$x$} \put(0,52){$y$} % \end{picture} \caption{$\mathscr{L}=\{x+2y,2x+y\}$} \label{figure:newta} % \end{figure} Then $w_\mathbf{0}(x^2f)=w_\mathbf{0}(x^3+x^2y)=w(x^3)=3/2$, so $LM(x^2f)=x^3$. On the other hand $w_\mathbf{0}(y^2f)=w_\mathbf{0}(xy^2+y^3)=w(y^3)=3/2$, hence $LM(y^2f)=y^3$. So $\prec$ is not a semigroup order. It follows from \cite{robbiano} that every semigroup ordering is a Newton ordering with $|\mathscr{L}|=1$ and ${\pmb{\delta}}=\mathbf{0}$. If the coefficients of every $l\in L$ are positive (negative), then $\prec$ is a local (global) ordering. Consider the set of monomials \begin{equation*} % \mathscr{M}_l =\{\bx^\balpha\in\mathscr{M}\mid w_{\pmb{\delta}}(\bx^\balpha) =l({\pmb{\alpha}}+{\pmb{\delta}})\}, \quad l\in\mathscr{L}. % \end{equation*} Clearly $\mathscr{M}_{\pmb{\delta}}(l)$ is a cone with respect to $\mathscr{M}(l)$: if $\bx^\balpha\in\mathscr{M}_{\pmb{\delta}}(l)$ and $\bx^\balpha\in\mathscr{M}(l)$, then $\bx^\balpha\bx^\bbeta\in\mathscr{M}_{\pmb{\delta}}(l)$. Let $R_{\pmb{\delta}}(l)$ be the $K$-algebra generated by the monomials in $\mathscr{M}_{\pmb{\delta}}(l)$. Then $R_{\pmb{\delta}}(l)$ is a $R(l)$-module. Since $\mathscr{L}$ is rational, $R(l)$ is noetherian and $R_{\pmb{\delta}}(l)$ is finitely generated since ${\pmb{\delta}}$ has rational coordinates. Now let $G={\{\fele}\}\subsetK[\bx]$ be a finite set of polynomials and $I\subseteqK[\bx]$ an ideal. Define $L(G)_{\pmb{\delta}}(l)=L(G)\cap\mathscr{M}_{\pmb{\delta}}(l)$ and $L(I)_{\pmb{\delta}}(l)=L(I)\cap\mathscr{M}_{\pmb{\delta}}(l)$. Let $R(G)_{\pmb{\delta}}(l)$ (resp.~$R(I)_{\pmb{\delta}}(l)$) be the $K$-algebra generated by the monomials in $L(G)_{\pmb{\delta}}(l)$ (resp.~$L(I)_{\pmb{\delta}}(l)$). \begin{example} % Let $\mathscr{L}=\{l_1,l_2,l_3\}=\{3x+3y,2x+6y,6x+2y\}$ and ${\pmb{\delta}}=(1,1)$. Then (figure \ref{figure:newtb}) % \begin{align*} % R(l_1) &= \mathbb{C}[xy,x^2y,x^3y,xy^2,xy^3],\quad \\ R(l_2) &= \mathbb{C}[xy^3,y]\quad\text{and}\quad \\ R(l_3) &= \mathbb{C}[x,x^3y]. % \end{align*} % \begin{figure}[H] % \centering \setlength{\unitlength}{1mm} \begin{picture}(55,55)(0,0) % \put(0,0){\epsfig{file=figure2.eps,height=5cm}} \put(52,0){$x$} \put(0,52){$y$} \put(35,35){$\mathscr{M}(l_1)$} \put(2,42){$\mathscr{M}(l_2)$} \put(38,3){$\mathscr{M}(l_3)$} % \end{picture} \caption{$\mathscr{L}=\{3x+3y,2x+6y,6x+2y\}$} \label{figure:newtb} % \end{figure} % The corresponding modules are (figure \ref{figure:newtc}): \begin{align*} % R_{\pmb{\delta}}(l_1) &= x^2R(l_1)\oplus x R(l_1)\oplus R(l_1)\oplus y R(l_1)\oplus y^2R(l_1),\\ R_{\pmb{\delta}}(l_2) &= y^2R(l_2)\quad\text{and}\quad\\ R_{\pmb{\delta}}(l_3) &= x^2R(l_3). % \end{align*} % \begin{figure}[H] % \centering \setlength{\unitlength}{1mm} \begin{picture}(55,55)(0,0) % \put(0,0){\epsfig{file=figure3.eps,height=5cm}} \put(52,0){$x$} \put(0,52){$y$} \put(35,35){$\mathscr{M}_{\pmb{\delta}}(l_1)$} \put(2,44){$\mathscr{M}_{\pmb{\delta}}(l_2)$} \put(40,3){$\mathscr{M}_{\pmb{\delta}}(l_3)$} % \end{picture} \caption{$\mathscr{L}=\{3x+3y,2x+6y,6x+2y\}$, ${\pmb{\delta}}=(1,1)$} \label{figure:newtc} % \end{figure} % \end{example} \begin{lemma}\label{lemma:sub} % $L(G)_{\pmb{\delta}}(l)$ and $L(I)_{\pmb{\delta}}(l)$ are cones with respect to $\mathscr{M}(l)$. $R(G)_{\pmb{\delta}}(l)$ and $R(I)_{\pmb{\delta}}(l)$ are finitely generated $R(l)$-submodules of $R_{\pmb{\delta}}(l)$. % \end{lemma} {\noindent\it Proof:\ } Let $f=\sum_{{\pmb{\alpha}}\in A} c_{{\pmb{\alpha}}}\bx^\balpha\inK[\bx]$ with $LM(f)=\mathbf{x}^{{\pmb{\alpha}}_0}\in\mathscr{M}(I)_{\pmb{\delta}}(l)$. For every $\bx^\bbeta\in\mathscr{M}(l)$ and every ${\pmb{\alpha}}\in A$ there exists a $l'\in\mathscr{L}$ such that \begin{align*} % w_{\pmb{\delta}}(\bx^\balpha\bx^\bbeta) &= l'({\pmb{\alpha}}+{\pmb{\beta}}+{\pmb{\delta}}) = l'({\pmb{\alpha}}+{\pmb{\delta}})+l'({\pmb{\beta}}) \\ & \geq l({\pmb{\alpha}}_0+{\pmb{\delta}})+l({\pmb{\beta}}) = w_{\pmb{\delta}}(\mathbf{x}^{{\pmb{\alpha}}_0}\bx^\bbeta). % \end{align*} so $LM(\bx^\bbeta f)=\bx^\bbeta LM(f)$. Therefore both $L(G)_{\pmb{\delta}}(l)$ and $L(I)_{\pmb{\delta}}(l)$ are cones with respect to $\mathscr{M}(l)$. This shows that $R(G)_{\pmb{\delta}}(l)$ and $R(I)_{\pmb{\delta}}(l)$ are indeed $R(l)$-submodules of $R_{\pmb{\delta}}(l)$. They are finitely generated since $R_{\pmb{\delta}}(l)$ is a noetherian module. $\square$\medskip\par \begin{proposition}\label{proposition:newtonnoether} % Let $\prec$ be a Newton ordering. Then $\prec$ is noetherian. % \end{proposition} {\noindent\it Proof:\ } Let $L_0\subseteq L_1\subseteq L_2\subseteq\ldots$ be an increasing sequence of lead monomial sets, i.e.~every $L_i$ is either $L(G)$ for a finite set of polynomials $G\subsetK[\bx]$ or $L(I)$ for an ideal $I\subseteqK[\bx]$. Let $l\in\mathscr{L}$. By lemma \ref{lemma:sub} the monomials of $L_i\cap\mathscr{M}_{\pmb{\delta}}(l)$ span over $K$ a $R(l)$-submodule $R_i$ of $R_{\pmb{\delta}}(l)$. Hence $R_0\subseteq R_1\subseteq R_2\subseteq\ldots$ is a increasing sequence of submodules. But $R_{\pmb{\delta}}(l)$ is noetherian, so this sequence gets stationary and so does the sequence $(L_i\cap\mathscr{M}_{\pmb{\delta}}(l))_{i\in\mathbb{N}}$. Since $|\mathscr{L}|<\infty$ also the sequence $(L_i)_{i\in\mathbb{N}}$ gets stationary. $\square$\medskip\par \begin{corollary} % Let $\prec$ be a Newton ordering. Then every ideal $I\subseteq K[\mathbf{x}]$ has a standard basis with respect to $\prec$. % \end{corollary} \begin{corollary} % Let $\prec$ be a normal Newton ordering. Then {\rm\bf NormalForm}\ terminates. % \end{corollary} Let $G={\{\fele}\}\subseteqK[t,\bx]$ be a finite set of homogeneous polynomials, let $I=(f_1,\ldots,f_r)$ and $f\in I$ homogeneous. We want to construct a reducing set for $(f,G)$. Let $S_l$ be a finite set of monomials which generates $R(\{f|_{t=1}\})_{\pmb{\delta}}(l)$ over $R(l)$, $l\in\mathscr{L}$ and let $S=\bigcup_{l\in\mathscr{L}}S_l$. For every $\bx^\bbeta\in S$ there exists a monomial $\bx^{\tilde{\bbeta}}\in\mathscr{M}$ such that $LM(\bx^{\tilde{\bbeta}} f|_{t=1})=\bx^\bbeta$. \begin{proposition}\label{proposition:newtonred} % $R=\{\bx^{\tilde{\bbeta}}\mid\bx^\bbeta\in S\}$ is a reducing set for $(f,G)$. % \end{proposition} {\noindent\it Proof:\ } Let $\bx^\balpha\in\mathscr{M}$ and let $t^{\gamma_0}\bx^\bgamma=LM(\bx^\balpha f)\in\mathscr{M}_{\pmb{\delta}}(l)$. We have to find a representation of $\bx^\balpha f$ as in definition \ref{definition:reducing_set}. Then there exist monomials $\bx^{\tilde{\bbeta}}\in R\cap\mathscr{M}_{\pmb{\delta}}(l)$ and $\bx^\btau\in\mathscr{M}(l)$ such that $\bx^\bgamma=\bx^\bbeta\bx^\btau$ and $\bx^\balpha=\bx^{\tilde{\bbeta}}\bx^\btau$. It follows from lemma \ref{lemma:normal_form} that there exists a homogeneous $g\inK[t,\bx]$ with $LM(g|_{t=1})=1$, $\xi_1,\ldots,\xi_k\in\mathscr{T}_t$ and $h_1,\ldots,h_k\in G\cup \{{\rm\bf NormalForm}(\bx^{\tilde{\bbeta}} f,G)\mid \bx^{\tilde{\bbeta}}\in R\}$ such that \begin{equation*} % g\bx^{\tilde{\bbeta}} f=\sum_{i=1}^k \xi_i h_i,\qquad LM(g\bx^{\tilde{\bbeta}} f)=LM(\xi_1h_1)\succ LM(\xi_ih_i),\quad 2\leq i\leq k. % \end{equation*} But then $g\bx^\balpha f=\bx^\btau g\bx^{\tilde{\bbeta}} f=\sum_{i=1}^k\bx^\btau\xi_i h_i$ with \begin{align*} % LM(g\bx^\balpha f) &= LM(\bx^\btau g\bx^{\tilde{\bbeta}} f) = \bx^\btau LM(g\bx^{\tilde{\bbeta}} f)\\ &= \bx^\btau LM(\xi_1h_1)= LM(\bx^\btau\xi_1 h_1)\\ &\succ LM(\bx^\btau\xi_i h_i), \quad 2\leq i\leq k. % \end{align*} This completes the proof. $\square$\medskip\par \begin{corollary} % Let $\prec$ be a normal Newton ordering. Then reducing sets do exist. % \end{corollary} \begin{corollary} % Let $\prec$ be a normal Newton ordering. Then {\bf StandardBasis} terminates. % \end{corollary} \section{Local orderings and zero dimensional ideals}\label{sect:local} Let $\prec$ be a local ordering and let $I\subseteq K[\mathbf{x}]$ be a zero dimensional ideal, i.e.~an ideal with $\dim_K(K[\mathbf{x}]/I)<\infty$. In this case $S_\loc^{-1}\Kx=K[\bx]_{(\mathbf{x})}=K[\bx]_{(x_1,\ldots,x_n)}$. We show that there always exists a finite standard basis of $I$. Moreover such a standard basis can be computed from a given standard basis with respect to another local semigroup ordering using only linear algebra. \begin{proposition}\label{proposition:always} % Every zero dimensional ideal $I\subseteq K[\mathbf{x}]$ has a standard basis with respect to any local order. % \end{proposition} {\noindent\it Proof:\ } Since $\dim I=0$ there exists a $m\in\mathbb{N}$ such that $(\mathbf{x})^m\subseteq I_{(\mathbf{x})}$. Then $L(I)\setminus (\mathbf{x})^m=\{\mathbf{x}^{{\pmb{\alpha}}_1}, \ldots,\mathbf{x}^{{\pmb{\alpha}}_k}\}$ is a finite set of monomials. There exist polynomials $f_1,\ldots,f_k\in\overline{I}$ with $LM(f_i)=\mathbf{x}^{{\pmb{\alpha}}_i}$, $1\leq i\leq k$. Clearly $\{f_1,\ldots,f_k\}$ together with all monomials of degree $m$ is a standard basis of $I$ with respect to $\prec$. $\square$\medskip\par For a local Newton ordering a standard basis can be computed with our variant of Buchbergers algorithm. For an arbitrary local ordering however it is not clear that this works. In particular, we do not know the corresponding reducing sets. One the other hand in the case of zero dimensional ideals and global semigroup orderings, one can do the following: given a standard basis with respect to a given ordering, the FGLM-algorithm \cite{fglm} computes a standard basis with respect to another ordering using only linear algebra. The key is that zero dimensional ideals admit finite defining systems. There are two reasons why the FGML-algorithm will not work in our case: Firstly, for a local ordering, a standard basis of $I$ does not necessarily generate $\overline{I}$ as a $K[\mathbf{x}]$-module. The second reason is that the FGLM-algorithm depends on the fact that the set of lead monomials has the structure of a monomial ideal, which is in general only true for semigroup orderings. Nevertheless it is possible to convert standard bases between different local orderings. Let $m\in\mathbb{N}$ such that $(\mathbf{x})^m\subseteq I_{(\mathbf{x})}$ and let $G={\{\fele}\}$ be a standard basis of $I$ with respect to $\prec$. Then $\overline{I}$ is an ideal in $K[\mathbf{x}]$ with $(\mathbf{x})^m\subseteq I_{(\mathbf{x})}$. The following lemma shows the relation between $G$ and $\overline{I}$. \begin{lemma} % $G$ together with all monomials of degree $m$ generates $\overline{I}$ as $K[\mathbf{x}]$-module. % \end{lemma} {\noindent\it Proof:\ } This follows immediately from $L(I)=L(\overline{I})$. $\square$\medskip\par \subsection*{The algorithm} First choose a monomial ideal $Z\subseteq\overline{I}$ which contains a power of the maximal ideal. The best choice for $Z$ would be the maximal monomial ideal contained in $\overline{I}$, but this ideal may be expensive to compute. Since $Z$ contains a power of the maximal ideal, $M\setminus Z$ is a finite set of monomials. Consider the following partition of $M\setminus Z$: \begin{align*} % N &= \{\mathbf{x}^{\pmb{\alpha}}\in M\setminus Z\mid \mathbf{x}^{\pmb{\alpha}}\not\in L(I)\},\\ H &= \{\mathbf{x}^{\pmb{\alpha}}\in M\setminus Z\mid \mathbf{x}^{\pmb{\alpha}}\in L(I)\}. % \end{align*} Let $H=\{\mathbf{x}^{{\pmb{\alpha}}_1},\ldots, \mathbf{x}^{{\pmb{\alpha}}_k}\}$ and $N=\{\mathbf{x}^{{\pmb{\alpha}}_{k+1}},\ldots, \mathbf{x}^{{\pmb{\alpha}}_{k+\mu}}\}$. There exist polynomials $h_1,\ldots,h_k\in\overline{I}$ such that $LT(h_i)=\mathbf{x}^{{\pmb{\alpha}}_i}$, $1\leq i\leq k$. These polynomials are straightforward to compute from $G$. The ideal $\overline{I}\subseteqK[\bx]$ is an infinite dimensional $K$-vector space with basis $\{h_1,\ldots,h_k\}\cup(\mathscr{M}\cap Z)$. So after some linear algebra we can assume $h_i-\mathbf{x}^{{\pmb{\alpha}}_i}\in K\left<N\right>$ (this simplification is just of cosmetic nature). Then \begin{equation*} % \left(\begin{array}{ccc|ccc} 1 & & 0 & \hphantom{0} & \hphantom{0} & \hphantom{0}\\ &\ddots & & \hphantom{0} & \ast & \hphantom{0}\\ 0 & & 1 & \hphantom{0} & \hphantom{0} & \hphantom{0} \end{array}\right) \begin{pmatrix} \mathbf{x}^{{\pmb{\alpha}}_1}\\ \vdots\\ \mathbf{x}^{{\pmb{\alpha}}_k}\\ \mathbf{x}^{{\pmb{\alpha}}_{k+1}}\\ \vdots\\ \mathbf{x}^{{\pmb{\alpha}}_{k+\mu}} \end{pmatrix} = \begin{pmatrix} h_1\\ \vdots\\ h_k \end{pmatrix}. % \end{equation*} Denote the $k\times(k+\mu)$ matrix on the left hand side by $A$. Now reorder the monomials $\mathbf{x}^{\alpha_1},\ldots,\mathbf{x}^{{\pmb{\alpha}}_{k+\mu}}$ with respect to a second local ordering $\prec_2$. This defines a matrix $\tilde{A}$ (which arises from $A$ by permuting columns) on which we perform Gauss elimination. Call the resulting matrix again $\tilde{A}$. Then there exists a permutation $\sigma\in S_{k+\mu}$ (operating on the columns of $\tilde{A}$) such that \begin{equation*} % \sigma(\tilde{A}) \begin{pmatrix} \mathbf{x}^{{\pmb{\alpha}}_{\sigma(1)}}\\ \vdots\\ \mathbf{x}^{{\pmb{\alpha}}_{\sigma(k)}}\\ \mathbf{x}^{{\pmb{\alpha}}_{\sigma(k+1)}}\\ \vdots\\ \mathbf{x}^{{\pmb{\alpha}}_{\sigma(k+\mu)}} \end{pmatrix} = \left(\begin{array}{ccc|ccc} 1 & & 0 & \hphantom{0} & \hphantom{0} & \hphantom{0}\\ &\ddots & & \hphantom{0} & \ast & \hphantom{0}\\ 0 & & 1 & \hphantom{0} & \hphantom{0} & \hphantom{0} \end{array}\right) \begin{pmatrix} \mathbf{x}^{{\pmb{\alpha}}_{\sigma(1)}}\\ \vdots\\ \mathbf{x}^{{\pmb{\alpha}}_{\sigma(k)}}\\ \mathbf{x}^{{\pmb{\alpha}}_{\sigma(k+1)}}\\ \vdots\\ \mathbf{x}^{{\pmb{\alpha}}_{\sigma(k+\mu)}} \end{pmatrix} = \begin{pmatrix} h'_1\\ \vdots\\ h'_k \end{pmatrix} % \end{equation*} with $LT(h_i')=\mathbf{x}^{{\pmb{\alpha}}_{\sigma(i)}}$ for $1\leq i\leq k$, thus $\mathbf{x}^{{\pmb{\alpha}}_{\sigma(1)}},\ldots\mathbf{x}^{{\pmb{\alpha}}_{\sigma(k)}}\in L_{\prec_2}(I)$. As a $K$-vector space $\overline{I}$ still has as basis $\{h'_1,\ldots,h'_k\}\cup (Z\cap\mathscr{M})$, so $\mathbf{x}^{{\pmb{\alpha}}_{\sigma(k+ 1 )}},\ldots, \mathbf{x}^{{\pmb{\alpha}}_{\sigma(k+\mu)}}\not\in L_{\prec_2}(I)$. Let $Z_{gen}$ be a finite set of generators of the monomial ideal $Z$. Then by construction $\{h'_1,\ldots,h'_k\}\cup Z_{gen}$ is a standard basis of $I$ with respect to $\prec_2$. Now we formulate the algorithm. So let $G$ be a standard basis of $I$ with respect to $\prec$. Assume that we are given the following procedures: \begin{description} % \item[${\rm\bf ComputeZgen}(G,\prec)$]\ \\ Input: $G$ is a standard basis of a zero dimensional ideal $I\subseteqK[\bx]$ with respect to the local ordering $\prec$.\\ Output: A reduced set of generators of a monomial ideal contained in $\overline{I}$ which contains a power of $(\mathbf{x})$. \item [${\rm\bf GaussEliminate}(A)$]\ \\ Input: $A$ is a $k\times(k+\mu)$ matrix over $K$ of rank $k$.\\ Output: The $k\times(k+\mu)$ matrix over $K$ which arises from $A$ by Gaussian elimination. % \end{description} \noindent Then the description of our algorithm is as follows. \begin{algorithm}\label{algorithm:convert}\ \newline\noindent % $G_2:={\rm\bf StandardBasisChange}(G,\prec,\prec_2)$\\ \hspace*{1cm}$Z_{gen}:={\rm\bf ComputeZgen}(G,\prec)$\\ \hspace*{1cm}$M:=\{\mathbf{x}^{\pmb{\alpha}}\in\mathscr{M}\mid\bx^\balpha\not\in (Z_{gen})\}$\\ \hspace*{1cm}$N:=\{\mathbf{x}^{\pmb{\alpha}}\in M\mid \bx^\balpha\not\in L(G)\}$\\ \hspace*{1cm}$H:=M\setminus N$\\ \hspace*{1cm}$F:=\emptyset$\\ \hspace*{1cm}FOR ALL $\bx^\balpha\in H$ DO\\ \hspace*{2cm}choose the first $f\in G$ with $LM(\eta f)=\bx^\balpha$ for some $\eta\in\mathscr{T}$\\ \hspace*{2cm}$F:=F\cup\{\eta f\}$\\ \hspace*{1cm}$A:=$ the matrix of coefficients of all $f\in F$ with respect to $M$\\ \hspace*{2cm}sorted according to $\prec_2$\\ \hspace*{1cm}$A:={\bf GaussEliminate}(A)$\\ \hspace*{1cm}$F:=$ the elements of the vector which arises from multiplying $A$ with\\ \hspace*{2cm}the column vector of the elements of $M$ sorted according to $\prec_2$\\ \hspace*{1cm}$G_2:=F\cup Z_{gen}$ % \end{algorithm} Altogether we have proved the \begin{proposition}\label{proposition:convert} % Let $G$ be a standard basis of the ideal $I\subseteqK[\bx]$ with respect to the local ordering $\prec$ and let $\prec_2$ be another local ordering. Then {\rm\bf StandardBasisChange}\ computes a standard basis of $I$ with respect to $\prec_2$. % \end{proposition} \begin{corollary} % For every ideal $I\subseteqK[\bx]$ and every local ordering $\prec$ % \begin{equation*} % \dim_KK[\bx]/\overline{I} = \left|\{\bx^\balpha\in\mathscr{M}\mid\bx^\balpha\not\in L(I)\}\right|. % \end{equation*} % \end{corollary} For degree orderings like $degrevlex^-$ the set of generators $Z_{gen}$ can be computed as follows: One computes the highest corner $\bx^\btau$ of $I$, i.e.~the smallest monomial not contained in $L(I)$. Then $\bx^\balpha\prec\bx^\btau$ implies $\bx^\balpha\in\overline{I}$. For degree orderings the set $\{\bx^\balpha\in\mathscr{M}\mid\bx^\balpha\succcurlyeq\bx^\btau\}$ is finite. Let $Z_1=\{\bx^\balpha\in\mathscr{M}\mid |{\pmb{\alpha}}|=|{\pmb{\tau}}|,\bx^\balpha\prec\bx^\btau\}$ and $Z_2=\{\bx^\balpha\in\mathscr{M}\mid |{\pmb{\alpha}}|=|{\pmb{\tau}}|+1,\bx^\balpha\not\in(Z_1)\}$. Then we may take $Z_{gen}=Z_1\cup Z_2$. This will also work (with a few modifications) in the case of a weighted degree ordering or a Newton ordering. \section{Applications}\label{sect:applications} Let $\prec$ be a normal Newton ordering, $I\subseteqK[\bx]$ an ideal and $G$ a standard basis of $I$ with respect to $\prec$. The Newton filtration on $K[\bx]$ induces a filtration on the quotient $K[\bx]/\overline{I}$. Let $Gr_\locK[\bx]/\overline{I}=\bigoplus_{s\in\mathbb{R}} (K[\bx]/\overline{I})_s/(K[\bx]/\overline{I})_{>s}$ be the associated graded module. Now assume that $\dim_K(K[\bx]/\overline{I})_s/(K[\bx]/\overline{I})_{>s}<\infty$ for all $s\in\mathbb{R}$ and let $P(t)$ be the Poincare series of $Gr_\locK[\bx]/\overline{I}$. The coefficients of this series can be computed as follows. \begin{proposition}\label{proposition:filtration} % Let $s\in\mathbb{R}$. The coefficient $c_s$ of $t^s$ in $P(t)$ is % \begin{equation*} % c_s = \dim_K(K[\bx]/\overline{I})_s/(K[\bx]/\overline{I})_{>s} = \left|\left\{\bx^\balpha\in M\mid \bx^\balpha\not\in L(I), w_{\pmb{\delta}}(\bx^\balpha)=s \right\}\right|. % \end{equation*} % \end{proposition} {\noindent\it Proof:\ } By definition the elements of $B=\left\{\bx^\balpha\in\mathscr{M}\mid\bx^\balpha\not\in L(I), w_{\pmb{\delta}}(\bx^\balpha)=s\right\}$ are linearly independent in the $K$-vector space $V=\left(K[\bx]\right/\overline{I})_s/\left(K[\bx]\right/\overline{I})_{>s}$. Now let $\bx^\balpha\in L(I)$ be the smallest monomial of Newton weight $w_{\pmb{\delta}}(\bx^\balpha)=s$. Then there exists a $f\in\overline{I}$ with $LM(f)=\bx^\balpha$. By construction $f-LT(f)\in K\left<B\right>\cupK[\bx]_{>s}$, so $\bx^\balpha\in K\left<B\right>$ in $V$. By induction this holds for all $\bx^\balpha\in L(I)$ of Newton weight $s$. $\square$\medskip\par \subsection{The spectrum of a hypersurface singularity} % Let $f\colon (\mathbb{C}^n,\mathbf{0})\rightarrow(\mathbb{C},0)$ be the germ of a holomorphic function with an isolated critical point at $\mathbf{0}$. Let $J_f= \left(\partial f/\partial x_1,\ldots, \partial f/\partial x_n\right)\subseteq\mathbb{C}\{\mathbf{x}\}$ be the Jacobi ideal of $f$. The Milnor number of $f$ % \begin{equation*} % \mu(f) =\dim_{\mathbb{C}}\left(\mathbb{C}\{\mathbf{x}\}/J_f\right) % \end{equation*} % is a topological invariant of $f$. Since $f$ has an isolated singularity $\mu(f)$ is finite. Then $f$ is $(\mu+1)$-determined, which means that $f$ does not change its analytical type if we forget about terms of order higher than $\mu+1$. So we can assume that $f$ is given by a polynomial. Moreover there exists a $m\in\mathbb{N}_0$ such that $(\mathbf{x})^m\subseteq J_f$, so $\mathbb{C}\{\mathbf{x}\}/J_f\simeq\mathbb{C}[\mathbf{x}]_{(\mathbf{x})}/J_f$. This means that we can compute the Milnor number of $f$ from a standard basis of $J_f$ with respect to a local ordering $\prec$ as % \begin{equation*} % \mu(f) =\left|\mathscr{M}\setminus L(J_f)\right|. % \end{equation*} % The spectrum of $f$ is an analytical invariant of $f$ which is finer than the Milnor number. It consists of $\mu=\mu(f)$ rational numbers $\{\alpha_1,\ldots,\alpha_\mu\}$ which often are written in the form $\sum_{i=1}^\mu t^{\alpha_i}$. These rational numbers are defined in terms of the mixed Hodge structure on the vanishing cohomology of the Milnor fibre of $f$ \cite{arnold}. In many cases the spectrum is determined by the Newton polygon of $f$ which is constructed as follows. If $f=\sum_{{\pmb{\alpha}}\in A}c_{\pmb{\alpha}}\bx^\balpha$ let $Or(\mathbf{x}^{\pmb{\alpha}})$ be the positive orthant in $\mathbb{R}^n$ centred at ${\pmb{\alpha}}\in\mathbb{N}_0^n\subset\mathbb{R}^n$, ${\pmb{\alpha}}\in A$. Then the Newton polygon $\Sigma_f$ of $f$ is the convex hull of the union $\bigcup_{{\pmb{\alpha}}\in A}Or(\mathbf{x}^{\pmb{\alpha}})$. Let $\Sigma_f^c$ be the set of compact faces of $\Sigma_f$. For every compact face $\sigma\in\Sigma_f^c$, let $f_\sigma =\sum_{{\pmb{\alpha}}\in A\cap\sigma}c_{\pmb{\alpha}}\bx^\balpha$ be the sum of the monomials of $f$ with support on $\sigma$. % \begin{definition} % $f$ has {\em nondegenerate principal part} if for every $\sigma\in\Sigma_f^c$ the polynomials $\partial f_\sigma/\partial x_1,\ldots, \partial f_\sigma/\partial x_n$ do not have a common root in $(\mathbb{C}^\ast)^n$. % \end{definition} % For every compact face $\sigma\in\Sigma_f^c$ there exists a unique linear form $l_\sigma\colon\mathbb{Q}^n\rightarrow\mathbb{Q}$ with $\left.l_\sigma\right|_\sigma\equiv 1$. Let $\mathscr{L}=\{l_\sigma\mid\sigma\in\Sigma_f^c\}$ and let ${\pmb{\delta}}=(1,\ldots,1)$. Consider the Newton filtration on $\mathbb{C}[\mathbf{x}]$ associated to $\mathscr{L}$ and ${\pmb{\delta}}$. This induces a filtration on $\mathbb{C}\{\mathbf{x}\}$. The following theorem is due to M.~Saito. % \begin{theorem}\label{theorem:saito} % \cite{saito,kv} If $f$ has nondegenerate principal part then the spectrum of $f$ coincides with the Poincare series of the artinian module $\mathbb{C}\{\mathbf{x}\}/J_f$ graded by the Newton filtration. % \end{theorem} % Now let $\prec$ be a Newton ordering for $\mathscr{L}$ and ${\pmb{\delta}}$. Clearly $\prec$ is local. Since we have assumed that $f$ is given by a polynomial, as a consequence of proposition \ref{proposition:filtration} and theorem \ref{theorem:saito} we get the % \begin{corollary}\label{corollary:spectrum} % If $f$ has nondegenerate principal part then the spectrum of $f$ consists of the $\mu$ rational numbers % \begin{equation*} % \{w_{\pmb{\delta}}\left(\bx^\balpha\right)\mid \bx^\balpha\not\in L(J_f)\}. % \end{equation*} % \end{corollary} % \begin{example} % Consider the surface singularity % \begin{align*} % f=&x^{12}+y^{12}+z^{12}+x^5y^5+x^5z^5+y^5z^5\\ &+xyz(x^2y^2+x^2z^2+y^2z^2)+x^2y^2z^2. % \end{align*} % The principal part of $f$ is nondegenerate and the Newton polygon of $f$ has twelve faces: % \begin{figure}[H] % \centering \setlength{\unitlength}{1mm} \begin{picture}(100,100)(0,0) % \put(10,100){\epsfig{file=figure4.ps,width=10cm,angle=270,% clip=, bbllx=50, bblly=200, bburx=550, bbury=700}} % \end{picture} \caption{Newton polygon of $f$} \label{figure:npoly} % \end{figure} % Using corollary \ref{corollary:spectrum} and algorithm \ref{algorithm:convert}, the spectrum of $f$ is % \begin{align*} % t^\frac{1}{2}&+ 3t^\frac{7}{12}+ 3t^\frac{2}{3}+ 6t^\frac{3}{4}+ 9t^\frac{5}{6}+ 9t^\frac{11}{12}+ 13t\\ +& 18t^\frac{13}{12}+ 18t^\frac{7}{6}+ 21t^\frac{5}{4}+ 24t^\frac{4}{3}+ 24t^\frac{17}{12}+ 25t^\frac{3}{2}\\ +& 24t^\frac{19}{12}+ 24t^\frac{5}{3}+ 21t^\frac{7}{4}+ 18t^\frac{11}{6}+ 18t^\frac{23}{12}+ 13t^2\\ +& 9t^\frac{25}{12}+ 9t^\frac{13}{6}+ 6t^\frac{9}{4}+ 3t^\frac{7}{3}+ 3t^\frac{29}{12}+ t^\frac{5}{2}. % \end{align*} % In particular $\mu(f)=323$ and $p_g(f)=44$. % \end{example} % \section{Concluding remarks} This work arose from the desire to have a program which computes the spectrum of an isolated hypersurface singularity. Our variant of the Buchberger algorithm has been successfully coded. However it turned out that it is much faster to compute a standard basis with respect to a semigroup ordering and to convert it to a standard basis with respect to the corresponding Newton ordering using ${\rm\bf StandardBasisChange}$. This not really a surprise: standard basis algorithms for semigroup orderings have been carefully optimised for years, see \cite{ggmnppss,macaulay}. The computation of the spectrum of an isolated hypersurface singularity with nondegenerate principal part is now implemented in the computer algebra program {\tt Singular} and will hopefully be available in further releases. For convenience, our web site \begin{center} % {\small\tt www.mathematik.uni-mainz.de/AlgebraischeGeometrie/Spectrum/index.shtml} % \end{center} offers an interface to this implementation. \subsection*{Acknowledgements} I would like to thank D.~van Straten for telling me about this problem, G.-M.~Greuel for valuable suggestions and H.~Sch\"onemann for his support during the implementation. G.-M.~Greuel, G.~Pfister and H.~Sch\"onemann gave me the opportunity to implement the computation of the spectrum of an isolated hypersurface singularity into {\tt Singular}.
\section{INTRODUCTION} Nowadays inflation is a widely accepted element of the early cosmology \cite {inf}. It gives the possibility of solving many of the shortcomings of the standard hot big bang model and provides the source for the early energy density fluctuations responsible of the large scale structure of the universe observed today. Although there are many models of inflation, the underlying physical ideas are well established. These are characterized by a period of ``slow roll'' evolution of a scalar field (called inflaton) toward the vacuum potential. During this period the field changes very slowly, so that the kinetic energy $\dot{\varphi}^{2}/2$ remains smaller than its potential energy $V(\varphi )$. The energy density associated to the scalar field acts as a ``cosmological constant'' term, allowing a period of quasi exponential expansion of the scale factor. When the period of inflation ends, the scalar field $\varphi $ start a phase of rapid coherent oscillations around the vacuum. Very recently (\cite{dm,lm}) it has been pointed out that inflation can persist during the coherent oscillations of the inflation field phase. This exciting result is possible when the inflaton potential verifies a simple constrain of curvature far from the core convex part, where the inflaton field can roll slowly. The efficiency of this phenomena could have important implications for \ GUT scale baryogenesis \cite{klr}. In fact, as suggested by Damour and Mukhanov (\cite{dm}),it can be expected that due to the increase of the oscillation frequency, there is the possibility to generate massive particles heavier than $\sim 10^{16}GeV$. In ref.\cite{dm} Damour and Mukhanov estimated the amount of inflation to be $\sim 10$ e-fold (powers of the scale factor). They argue that this effect can be more efficient than the parametric resonance effect \cite{kls} for the amplification of cosmological perturbations \cite{fi-bran}.\ In ref.\cite {lm} Liddle and Mazumdar showed that Mukhanov et al. overestimated the number of e-fold because they have used a slow-roll definition of this object. In their paper, Liddle and Mazumdar found an analytical expression for the number of e-fold of inflation using the appropriate definition finding a number of $\sim 3$ e-fold concluding that this effect is not very efficient. The study of adiabatic perturbations in this phase has been made by Taruya \cite{taru}. He found a poor amplification in the case of a single scalar field model but anticipated an enormous amplification for multi-field systems. In this letter we review the problem. In particular we find that the analytical expressions used to compare with the numerical estimation are not well defined in the $q\sim 0$ region and propose a way to correct these analytical estimations. Furthermore, with this result we study the evolution of the scalar field finding total agreement with the conclusions of ref.\cite {lm} for $q>0.2$, but a remarkable different result for small $q$. For this region, the initial conditions are very important. We find that $q\sim 0$ gives the leading contribution for oscillating inflation and the dominant part in the amplification of the fluctuations. The letter is organized as follow; first we describe briefly the Damour-Mukhanov model. Then, we make some comments about the initial conditions for this phenomenon and later we propose an improved expression, valid for the leading region of $q$, which is our main contribution. \section{BASIC EQUATIONS} Now we shall restrict ourselves to models of inflation driven by a single scalar field. The equations are \begin{equation} \ddot{\varphi}+3H\dot{\varphi}+V,_{\varphi }=0, \label{ec1} \end{equation} \begin{equation} H^{2}=\kappa ^{2}(\frac{1}{2}\dot{\varphi}^{2}+V), \label{ec2} \end{equation} Here $H=\dot{a}/a$ is the Hubble parameter, $a$ is the scale factor of the universe and $\kappa ^{2}=8\pi /3M_{p}^{2}$ with $M_{p}=1.2\cdot 10^{19}Gev$ the Planck mass. During the oscillatory phase of $\varphi $ we have two time scales; the inverse of the frequency $\omega ^{-1}$ of oscillations of $% \varphi $ and the inverse of the rate of expansion $H^{-1}$. If the limit $% \omega \gg H$ is taken we can neglect terms proportional to $H$ in the equations. So from (1) we can integrate to obtain, \begin{equation} \rho =\frac{1}{2}\dot{\varphi}^{2}+V=cte=V_{m}, \label{ta1} \end{equation} where $V_{m}=V(\varphi _{m})$ is the maximum value of $V(\varphi )$ in each oscillation when the field reaches the maximum value $\varphi _{m}$. From this relation we obtain the period of a single oscillation, $% T=4\int_{0}^{\varphi _{m}}d\varphi \left[ 2(V_{m}-V(\varphi )\right] ^{\frac{% 1}{2}}.$When $\omega \ll H$ we can define an adiabatic average index $\gamma $ by $\gamma =\left\langle (\rho +p)/\rho \right\rangle ,$where the bracket means $\left\langle ...\right\rangle =T^{-1}\int_{0}^{T}...dt$. Equations (1,2) can be re-written in the fluid form \begin{equation} \dot{\rho}=-3H(p+\rho ), \label{ec3} \end{equation} \begin{equation} \frac{\ddot{a}}{a}=-\frac{1}{3}(\rho +3p), \label{ec4} \end{equation} then from the definition of $\gamma $ and eqns.(\ref{ec3},\ref{ec4}) we have several ways to compute the adiabatic index \begin{equation} \gamma =\frac{\left\langle \dot{\varphi}^{2}\right\rangle }{V_{m}}=\frac{% \left\langle \varphi V,_{\varphi }\right\rangle }{V_{m}}=2(1-\frac{% \left\langle V\right\rangle }{V_{m}}). \label{g} \end{equation} Because $p=(\gamma -1)\rho $ and (\ref{ec4}) we have a superluminal expansion $\ddot{a}>0$ when $\gamma <2/3$. From the last two relations in eqn.(\ref{g}) the inequality $\gamma <2/3$ leads to \begin{equation} \left\langle V-\varphi V,_{\varphi }\right\rangle >0. \label{rm} \end{equation} \section{THE DAMOUR-MUKHANOV MODEL} Until now everything has been done for an arbitrary potential, but from now on we shall consider the potential \begin{equation} V(\varphi )=\frac{A}{q}\left[ \left( \frac{\varphi ^{2}}{\varphi _{c}^{2}}% +1\right) ^{q/2}-1\right] , \label{pot} \end{equation} where $q$ is a dimensionless parameter, $A=[$mass$]^{4}$ is a constant and $% \varphi _{c}=[$mass$]$ determines the size of the convex core of $V(\varphi ) $. We assume for a while that $\varphi _{c}$ marks the end of {\it % oscillating inflation}. The analysis made in ref.\cite{dm} works well far from the core of the potential. Further, the limit $\varphi \gg \varphi _{c}$ of eqn.(\ref{pot}) was written as: \begin{equation} V(\varphi )\simeq \frac{A}{q}\left( \frac{\varphi }{\varphi _{c}}\right) ^{q}. \label{pot1} \end{equation} In this case, the adiabatic index can be computed exactly given \cite{mt} by \begin{equation} \gamma =\frac{2q}{q+2}, \label{gama} \end{equation} so, from the inequality $\gamma <2/3$ we note that to hold inflation during the oscillatory phase we must have $q<1$. By using eqn.(\ref{gama}) in eqn.(% \ref{ec3}) we obtain $\dot{\rho}=-3H\gamma \rho $ and together with eqn.(\ref {ec2}) we have \begin{equation} a\propto t^{2/3\gamma }=t^{(q+2)/3q}, \label{ee1} \end{equation} \begin{equation} \varphi _{m}\propto t^{-2/q}\propto a^{-6/(q+2)}, \label{ee2} \end{equation} \begin{equation} \rho =V(\varphi _{m})\propto t^{-2}\propto a^{-6q/(q+2)}, \label{ee3} \end{equation} where $\varphi _{m}$ is the amplitude of the oscillations, $\varphi _{c}<\varphi _{m}<\varphi _{s}$ and $\varphi _{s}$ is a typical value of $% \varphi $ at the end of slow-roll inflation and the beginning of oscillating inflation. To compute the number of e-fold of inflation during oscillating inflation we can not use the standard expression $N=\ln (a_{f}/a_{i})$, appropriate for the slow-roll stage, but the improved expression proposed in ref.\cite{lm} \begin{equation} \tilde{N}=\ln \frac{a_{f}H_{f}}{a_{i}H_{i}}, \label{ne} \end{equation} because in each oscillation, while the field spends time in the core region, the universe continue their expansion so $H$ can vary. Then from (\ref{ec2}) and (\ref{ee3}) $H\propto a^{-3q/(q+2)}$ the product $aH\propto \varphi _{m}^{(1-q)/3}$ and from (\ref{ne}) we obtain \begin{equation} \tilde{N}\simeq \frac{1-q}{3}\left[ \ln \frac{qM_{p}}{\varphi _{c}}-2\right] , \label{ne2} \end{equation} where we have used $\varphi _{s}\sim qM_{p}/\sqrt{16}\pi $. In \cite{lm} the numerical curves for $\varphi _{c}=10^{-6}M_{p}$ show that $\tilde{N}% \lesssim 3$. Using the analytical expression (\ref{ne2}) we do not find agreement for small values of $q$. However there is not a compelling reason to believe in (\ref{ne2}) for small $q$. \section{THE SMALL q-REGION} To study the small $q$ region, we must use the correct limit $q\rightarrow 0$ of (\ref{pot}) which leads to \begin{equation} V(\varphi )\simeq \frac{A}{2}\ln \left[ \left( \frac{\varphi }{\varphi _{c}}% \right) ^{2}+1\right] , \label{pot2} \end{equation} so, if now we take the limit $\varphi \gg \varphi _{c}$ we obtain the logarithmic potential $V(\varphi )\simeq A\ln (\varphi /\varphi _{c})$. A very important fact to note from (\ref{pot1}) is that the limit $% q\rightarrow 0$ does not exist. Of course, the expression (\ref{pot1}) is wrong around the $q\sim 0$ region and the expressions derived from this are ill-defined. But, some work has been done in this regard \cite{dm}. For the logarithmic potential the adiabatic index is $\gamma =1/\ln (\varphi _{m}/\varphi _{c})$, so from (\ref{ec3}) and (\ref{ec2}) we obtain \begin{equation} a(t)\propto \exp \left[ -\frac{A}{2}(t_{end}-t)^{2}\right] , \label{ta3} \end{equation} but this form does not permit us to write an explicit expression for $\tilde{% N}$ (see ref. \cite{lm}). Let us make some comments about this result. Because $\gamma =1/\ln (\varphi _{m}/\varphi _{c})$ from (\ref{ec3}) we obtain $\dot{\rho}=-3HA$, then $a\sim (\varphi _{m}/\varphi _{c})^{-1/3}$. Moreover, from (\ref{ec2}) we have $H\propto \rho ^{1/2}\propto (\ln (\varphi _{m}/\varphi _{c}))^{1/2}$, then to compute $\tilde{N}$ we should evaluate the factor $(\ln (\varphi _{m}/\varphi _{c}))^{1/2}(\varphi _{m}/\varphi _{c})^{-1/3}$ at the extremes $\varphi _{s}$ and $\varphi _{c}$% , but this is not possible. The problem arises when $\varphi _{m}$ is chosen close to $\varphi _{c}$ in a expression valid for $\varphi \gg \varphi _{c}$% . Because (\ref{pot1}) is valid for $\varphi \gg \varphi _{c}$ too, the same problem should appear in the calculation of $\tilde{N}$. In fact this is the case but, to see that, we must include the constant term $A/q$ in (\ref{pot1}% ) or equivalently, write the limit correctly. When we use the potential \begin{equation} V(\varphi )\simeq \frac{A}{q}\left[ \left( \frac{\varphi }{\varphi _{c}}% \right) ^{q}-1\right] , \label{pot3} \end{equation} the adiabatic index become time-dependent, satisfying the equation \begin{equation} \gamma \rho =\frac{2q}{q+2}(\rho +\frac{A}{q}), \label{ta4} \end{equation} where $\rho (t)=V(\varphi _{m}(t))=A[(\varphi _{m}(t)/\varphi _{c})^{q}-1]/q. $ Replacing this in (\ref{ec3}) we obtain the same behavior as in eqn.(\ref{ee2}). But when we calculate $H$ we obtain $H\propto \rho ^{1/2}\propto q^{-1/2}[(\varphi _{m}/\varphi _{c})^{q}-1]^{1/2}$, which is not well defined at the point $\varphi _{m}=\varphi _{c}$. The same happens to the e-fold number, as mentioned before. The authors of \cite{lm} found $\varphi _{s}\sim qM_{p}/\sqrt{16}\pi $ using the potential (\ref{pot1}). From this result they found a decrease of $% \tilde{N}$ at decreasing values of $q$. This seems very strange because, for smaller values of $q$ we obtain flatter potentials at $\varphi \gg \varphi _{c}$, then $\varphi $ rolls slowly most of the time increasing the amount of inflation. Furthermore, we are not safe of how they set the initial conditions for $\varphi $ in their numerical analysis. In general, the initial and final field configuration depends on the potential. For $q$ close to cero, $\varphi _{s}$ is not proportional to $q$. In fact, we know that $\varphi _{s}$ came from the saturation of the slow-roll inequality $\left| V^{\prime }/V\right| <\sqrt{48\pi }/M_{p}$ using the expression (\ref{pot3}) for the potential \begin{equation} \left( \frac{\varphi _{s}}{\varphi _{c}}\right) ^{q-1}=\frac{\sqrt{48\pi }% \varphi _{c}}{qM_{p}}\left[ \left( \frac{\varphi _{s}}{\varphi _{c}}\right) ^{q}-1\right] . \label{recur} \end{equation} for the $\varphi \gg \varphi _{c}$ region. In Figure 1 we see the behavior of $\varphi _{s}$ in terms of $q$ given by eqn.(\ref{recur}). Of course, in the large $q$-region both curves agree. In order to illustrate this point and compare with ref.\cite{lm} we use its expression given by eqn.(\ref{ne}) but instead of using $\varphi _{s}\sim qM_{p}/\sqrt{16}\pi $ we use the numerical values of $\varphi _{s}$ obtained from eqn.(\ref{recur}). The results are plotted in Figure 2. In the small $q$-region the field $\varphi _{s}$ grows preventing the fall of $\tilde{N}$ predicted in ref.\cite{lm}. Moreover, as we have anticipated before, the value of the field at the end of oscillating inflation will have a $q$-dependence too. We know from ref. \cite{dm} that the intercept $U(\varphi )=V(\varphi )-\varphi V_{,\varphi }$% , must be positive to hold oscillating inflation. Let us define $\varphi _{f} $ to be the value of the inflaton field $\varphi $ at which $U(\varphi _{f})=0$. This condition represents the end of inflation due to oscillation. We need thus to compare $\varphi _{f}$ for different values of $q$. If we take the potential (\ref{pot3}) and define $x=\varphi /\varphi _{c}$, we obtain \begin{equation} U(x)=\frac{A}{q}\left[ x^{q}\left( 1-q\right) -1\right] . \label{ta6} \end{equation} From this equation we can extract a explicit expression for $\varphi _{f}$. If we impose $U(x_{f})=0$ we obtain the value of the scalar inflaton field at the end of this phase \begin{equation} \varphi _{f}=\varphi _{c}\left( 1-q\right) ^{-1/q}. \label{fif} \end{equation} Using the improved expression eqn.(\ref{ne}) for the e-fold number (ref.\cite {lm}), but inserting the corrected values for $\varphi _{s}$ and $\varphi _{f}$ given by eqns. (\ref{recur}) and (\ref{fif}) we obtain \begin{equation} \tilde{N}\simeq \ln \left\{ \left( \frac{\varphi _{s}}{\varphi _{f}}\right) ^{(2+q)/6}\left[ \frac{(\varphi _{f}/\varphi _{c})^{q}-1}{(\varphi _{s}/\varphi _{c})^{q}-1}\right] ^{1/2}\right\} . \label{ta7} \end{equation} The corrected value for $\varphi _{f}$ leads to a even smaller amount of inflation when comparing with the value obtained in ref.\cite{lm}. In figure 3, we plot eqn.(\ref{ta7}) and show the behavior of both effects combined. Because $\varphi _{f}$ is greater than $\varphi _{c}$, the amount of inflation is smaller than one obtained by Liddle et al.\cite{lm} in the whole range of $q\in (0,1)$. Moreover, the correct values of $\varphi _{s}$ produce a positive contribution to $\tilde{N}$ in the small $q$-range, which avoids the fall of $\tilde{N}$, as $q$ goes to smaller values, predicted by Liddle et. al.\cite{lm}. Again, it is not possible to show the whole range of $q$ because $q>\varphi _{c}/\varphi $ (see comments below eqn.(\ref{recur}% )). Because oscillating inflation adds e-folds of inflation after the slow-roll regime, where the observed perturbations are generated, is possible that this additional period of inflation could push perturbations to observable scales. In order to obtain the required amplitude of density perturbations and without imposing unphysical constraint on the potential, we should compute the density perturbation spectrum for the model being studied. In reference \cite{lm} an expression for this object was derived \[ \delta _{H}^{2}=\frac{512\pi }{75}\frac{A}{q^{3}M_{p}^{6}}\frac{\varphi ^{q+2}}{\varphi _{c}^{q}}. \] However, this expression is not well defined close enough to zero. Using the primordial density perturbation spectrum $\delta _{H}$ as was define in \cite {lm} we obtain for eqn.(\ref{pot3}) \begin{equation} \delta _{H}^{2}=\frac{512\pi }{75}\frac{A}{q^{3}M_{p}^{6}}\left\{ \frac{% \left[ \left( \varphi /\varphi _{c}\right) ^{q}-1\right] ^{3}}{\left( \varphi /\varphi _{c}\right) ^{2q}}\right\} , \label{pertur} \end{equation} which is well defined even for the small values of $q$ : \begin{equation} \delta _{H}^{2}\approx \frac{512\pi }{75}\frac{A}{M_{p}^{6}}\ln ^{3}\left( \frac{\varphi }{\varphi _{c}}\right) . \label{pertur2} \end{equation} Because the COBE satellite require $\delta _{H}\approx 2\cdot 10^{-5}$, in the $q=0$ case the amplitude of the potential for $\varphi _{c}=10^{-6}M_{p}$ gives $A^{1/4}\sim 2\cdot 10^{-3}M_{p}$, which is a typical number for inflationary models. \section{SUMMARY} As a summary, we have made some corrections about how to compute the e-fold number, which accounts for the amount of inflation during the oscillatory phase. In particular we note that previous studies are not accurate because they are not valid close to the core of the potential $\varphi \sim \varphi _{c}$. Thus, we make the analysis for the small $q$ region of the potential, which has not been considered until now. Finally we find that, in order to extract the correct amount of inflation during this phase, a very careful definition of initial and final field configuration is needed. Our results show that near $q\sim 0$ the e-fold number is maximal but it is still not enough to be more efficient than the parametric resonant effect discussed in \cite{kls}. \section*{Acknowledgments} The authors want to thanks Sergio del Campo for helpful discussions. V. C\'{a}rdenas wants to thanks CONICYT for support through a scholarship. G. Palma was supported in part through Proyect FONDECYT 1980608 and Proyect DICYT 049631PA.
\section{Introduction} \label{sec:1} Consider the randomly forced Burgers equation: \begin{veqnarray} \label{eq:1.1} u_t+u u_x =\nu u_{xx} +f, \end{veqnarray} where $f(x,t)$ is a zero-mean, Gaussian, statistically homogeneous, and white-in-time random process with covariance \begin{veqnarray} \nonumber \< f(x,t) f(y,s)\> = 2 B(x-y)\delta(t-s), \end{veqnarray} and $B(x)$ is a smooth function. In the language of stochastic differential equation (\ref{eq:1.1}) must be interpreted as \begin{veqnarray} \nonumber du =(-u u_x +\nu u_{xx}) dt+dW(x,t), \end{veqnarray} where $W(x,t)$ is a Wiener process with covariance \begin{veqnarray} \nonumber \< W(x,t)W(y,s)\>=2B(x-y)\min(t,s). \end{veqnarray} We will be interested in the statistical behavior of the stationary states (invariant measures) of (\ref{eq:1.1}) if they exist, or the transient states with possibly random initial data. There are two main reasons for considering this problem. The first is that (\ref{eq:1.1}) and its multi-dimensional version are among the simplest nonlinear models in non-equilibrium statistical mechanics. As such they serve as qualitative models for a wide variety of problems including charge density waves \cite{feig83}, vortex lines in high temperature super-conductors \cite{blfege94}, dislocations in disordered solids and kinetic roughening of interfaces in epitaxial growth \cite{krsp92}, formation of large-scale structures in the universe \cite{shze89,vedufr94}, etc. The connection between these problems and (\ref{eq:1.1}) can be understood as follows. Consider an elastic string in a random potential $V(x,s)$. The string is assumed to be {\em directed} in the sense that there is a time-like direction, assumed to be~$s$, such that the configuration of the string is a graph over the $s$-axis. Let $Z(x,t)$ be the partition function for the configurations of the string in the interval $0\le s\le t$, pinned at position~$x$ at time~$t$: \begin{veqnarray} \nonumber Z(x,t) = \left.\left\< {\rm e}^{-\beta \int_0^t ds\ V(w(s),s)} \right|w(t)=x\right\>, \end{veqnarray} where $\<\cdot|w(t)=x\>$ denotes the expectation over all Brownian paths $w(\cdot)$ such that $w(t)=x$, $\beta=1/kT$, $k$ is the Boltzmann constant, $T$ is the temperature. The free energy, $\varphi(x,t)=\ln Z(x,t)$ then satisfies \begin{veqnarray} \nonumber \varphi_t +{\t\frac{1}{2}} |\nabla \varphi|^2 = \nu \Delta \varphi +V, \end{veqnarray} where $\nu=kT$. This is the well-known Kardar-Parisi-Zhang equation \cite{kapazh86}. In one dimension if we let $u=\varphi_x$, $f=V_x$, we obtain (\ref{eq:1.1}). The second reason for studying (\ref{eq:1.1}) is in some sense a technical one. For a long time, (\ref{eq:1.1}) has served as the benchmark for field-theoretic techniques such as the direct interaction approximation or the renormalization group methods, developed for solving the problem of hydrodynamic turbulence. This role of (\ref{eq:1.1}) is made more evident by the recent flourish of activities introducing fairly sophisticated techniques in field theory to hydrodynamics \cite{pol95,bomepa95,gumi96}. In this context (\ref{eq:1.1}) is often referred to as Burgers turbulence. Since the phenomenology of the so-called Burgers turbulence is far simpler than that of real turbulence, one hopes that exact results can be obtained which can then be used to benchmark the methods. However so far our experience has proved otherwise: The problem of Burgers turbulence is complicated enough that a wide variety of predictions have been made as a consequence of the wide variety of techniques used \cite{ave95b,ave95a,avrye94,bafako97,% bol97,bol98,bome96,chya95a,chya95b,chya96,ekhma97,eva99a,got94,gokr93,gokr98,% gumi96,gusiau97,kra99,pol95,ryav99,sin92,sin96,vedufr94}. The main purpose of the present paper is to clarify this situation and to obtain exact results that are expected for Burgers turbulence. Along the way we will also develop some technical aspects that we believe will be useful for other problems. The main issues that interest us are the scaling of the structure functions and the asymptotic behavior of the probability density functions (PDF) in the inviscid limit. The former is well-understood heuristically but we will derive the results from self-consistent asymptotics on the master equation. The latter is at the moment very controversial, and we hope to settle the controversy by deriving exact results on the asymptotic behavior for the PDFs. From a technical point of view we insist on working with the master equation and making no closure assumptions. The unclosed term is expressed in terms of the statistics of singular dissipative structures of the field, here the shocks. We then derive a master equation for the statistics of the environment of the shocks by relating them to the singular structures on the shocks, namely the points of shock creation and collisions. These are then amenable to local analysis. In this way we achieve closure through dimension reduction. We then extract information on the asymptotic behavior of PDFs using realizability constraints and self-consistent asymptotics. We certainly hope that this philosophy will be useful for other problems. One main issue that will be addressed in this paper is the behavior of the PDF of the velocity gradient. Assuming statistical homogeneity, let $Q^\nu(\xi,t)$ be the PDF of $\xi=u_x$. $Q^\nu$ satisfies \begin{veqnarray} \nonumber Q^\nu_t=\xi Q^\nu +\bigl(\xi^2 Q^\nu\bigr)_\xi +B_1 Q^\nu_{\xi\xi} -\nu\bigl( \< \xi_{xx}|\xi\> Q^\nu\bigr)_\xi, \end{veqnarray} where $B_1=-B_{xx}(0)$, $\<\xi_{xx}|\xi\>$ is the average of $\xi_{xx}$ conditional on $\xi$. This equation is unclosed since the explicit form of the last term, representing the effect of the dissipation, is unknown. We are interested in $Q^\nu$ at the inviscid limit: \begin{veqnarray} \nonumber \underline{Q}(\xi,t)=\lim_{\nu\to0} Q^\nu(\xi,t). \end{veqnarray} In order to derive an equation for $\underline{Q}$, one needs to evaluate \begin{veqnarray} \nonumber F(\xi,t)= -\lim_{\nu\to0} \nu\bigl( \< \xi_{xx}|\xi\> Q^\nu\bigr)_\xi. \end{veqnarray} This is where the difficulty arises. Remembering that $u_x(x,t)=\lim_{{\it\Delta}x} \def\du{{\it \Delta}u} \def\dv{{\it \Delta}v\to0}(u(x+{\it\Delta}x} \def\du{{\it \Delta}u} \def\dv{{\it \Delta}v,t)-u(x,t))/{\it\Delta}x} \def\du{{\it \Delta}u} \def\dv{{\it \Delta}v$, the procedure outlined above pertains to the process of first taking the limit as ${\it\Delta}x} \def\du{{\it \Delta}u} \def\dv{{\it \Delta}v\to0$, then the limit as $\nu\to0$. It is natural to consider also the other situation when the limit $\nu\to0$ is taken first. In this case the limiting form of (\ref{eq:1.1}), (notice that we now write the nonlinear term in a conservative form), \begin{veqnarray} \label{eq:1.2} u_t+{\t\frac{1}{2}}\bigl( u^2\bigr)_x = f, \end{veqnarray} has to be interpreted in a weak sense by requiring \begin{veqnarray} \label{eq:1.3} \hbox{$\d\int\!\!\!\!\int$} dxdt\ \Bigl\{ u\varphi_t +{\t\frac{1}{2}}u^2\varphi_x + f\varphi\Bigr\} =0, \end{veqnarray} for all compactly supported smooth functions $\varphi$. The solutions $u$ satisfying (\ref{eq:1.3}) are called weak solutions. In order to ensure a well-defined dynamics for (\ref{eq:1.2}), i.e. existence and uniqueness of solutions with given initial data, an additional entropy condition has to be imposed on weak solutions. This amounts to requiring \begin{veqnarray} \label{eq:1.4} u(x{\scr+},t)\le u(x{\scr-},t), \end{veqnarray} for all $(x,t)$. The entropy condition (\ref{eq:1.4}) is the effect of the viscous term in the inviscid limit. There is a huge mathematical literature on (\ref{eq:1.2}) for the deterministic case. Standard references are \cite{lax57,lax73,smo83}. The random case was studied recently in the paper by E, Khanin, Mazel and Sinai \cite{ekhma99}. The first step in the present paper is to derive master equations for single and multi-point statistics of $u$ satisfying (\ref{eq:1.2}). In particular, we derive an equation for the PDF of $\eta(x,y,t)=(u(x+y)-u(x,t))/y$, $Q^\delta(\eta,x,t)$. We are interested in \begin{veqnarray} \nonumber Q(\xi,t)=\lim_{x\to0}Q^\delta(\xi,x,t). \end{veqnarray} One natural question is whether \begin{veqnarray} \label{eq:1.4.1} Q=\underline Q. \end{veqnarray} We will present very strong argument that (\ref{eq:1.4.1}) holds for generic initial data and for the type of forces described after (\ref{eq:1.1}). Some of our results also apply to $Q$ for the more general case when it is possibly different from $\underline Q$. The issue now reduces to the evaluation or approximation of $F(\xi,t)$. Several different proposals have been made, each leads at statistical steady state ($Q_t=0$) to an asymptotic expression of the form \begin{veqnarray} \nonumber Q(\xi) \sim \left\{ \begin{array}{ll} {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle C_-|\xi|^{-\alpha}}& {\rm as} \ \ \xi\to-\infty,\\[6pt] {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle C_+\xi^\beta {\rm e}^{-\xi^3/(3B_1)}\qquad}& {\rm as} \ \ \xi\to+\infty, \end{array}\right. \end{veqnarray} but with a variety of values for the exponents $\alpha$ and $\beta$ (here the $C_\pm$'s are constants). By invoking the operator product expansion, Polyakov \cite{pol95} suggested that $F=aQ+b\xi Q$, with $a=0$ and $b=-1/2$. This leads to $\alpha=5/2$ and $\beta=1/2$. Boldyrev \cite{bol97,bol98} considered the same closure with $-1\leq b\leq 0$, which gives $2\leq \alpha\leq 3$ and $\beta=1+b$. Based on heuristic arguments, Bouchaud and M{\'e}zard \cite{bome96} introduced a Langevin equation for the local slope of the velocity, which gives $2\le \alpha\le 3$, $\beta=0$. The instanton analysis \cite{bafako97,fakole96,gumi96} predicts the right tail of $Q$ without giving a precise value for $\beta$, and it does not give any specific prediction for the left tail. E~{\it et~al.} \cite{ekhma97} made a geometrical evaluation of the effect of $F$, based on the observation that large negative gradients are generated near shock creation. Their analysis gives a rigorous upper-bound for $\alpha$: $\alpha \le 7/2$. In~\cite{ekhma97}, it was claimed that this bound is actually reached, i.e., $\alpha=7/2$. Finally Gotoh and Kraichnan \cite{gokr98} argued that the viscous term is negligible to leading order for large ${|\xi|}$, i.e. $F\approx 0$ for $|\xi|\gg B_1^{1/3}$, giving rise to $\alpha=3$ and $\beta=1$. In this paper, we will first give a more detailed proof of the bound \begin{veqnarray} \nonumber \alpha>3, \end{veqnarray} announced in \cite{eva99a}. We will then show that \begin{veqnarray} \nonumber \alpha=7/2, \end{veqnarray} by studying the master equation for the environment of the shocks. An important reason behind the success of this program is that local behavior near the most singular structures (here the points of shock creation) is understood. Points of shock creation were isolated in \cite{ekhma99} as a mechanism for obtaining large negative values of $u_x$ and resulted in the prediction that $\alpha=7/2$. From this point of view, the present paper provides the missing step establishing the fact that points of shock creation provide the leading order contribution to the left tail of $Q$. For the most part, our working assumption will be that solutions of (\ref{eq:1.2}) are piecewise smooth. Shock path are smooth except at the points of collision. In particular, shocks are created at zero amplitude and the shock strength adds up at collision. For transient states these statements follow from standard results for the deterministic case under appropriate conditions for the initial data \cite{liu99}. For stationary states, these results are of the type established in \cite{ekhma99} under a non-degeneracy condition for the forcing. Before ending this introduction, we make some remarks about notations and nomenclature. In analogy with fluid mechanics $u$ will be referred to as the velocity field. We will denote the multi-point PDFs of $u(\cdot,t)$ as $Z^\nu(u_1,x_1,\ldots,u_n,x_n,t)$, i.e. \begin{veqnarray} \nonumber &&\hbox{Prob} (a_1\le u(x_1,t)\le b_1,\ldots,a_n\le u(x_n,t)\le b_n)\\ &=& \int_{a_1}^{b_1}\!\!du_1\cdots \int_{a_n}^{b_n}\!\!du_n\ Z^\nu(u_1,x_1,\ldots,u_n,x_n,t). \end{veqnarray} The superscript $\nu$ refers to the viscous case, $\nu>0$. In the inviscid limit we will denote the multi-point PDFs of $u(\cdot,t)$ by $Z(u_1,x_1\ldots,u_n,x_n,t)$. Statistical stationary values will be denoted by the subscript $\infty$, e.g. $Z^\nu_\infty$ or $Z_\infty$. We reserve the special notations $R^\nu(u,x,t)=Z^\nu(u,x,t)$ and $R(u,x,t)=Z(u,x,t)$ for the one-point PDF of $u(x,t)$. $Q^\nu(\xi,t)$ denotes the PDF of $\xi(x,t)$, and its inviscid limit is $Q(\xi,t)$. Statistical symmetries, or equality in law, will be denoted by $\stackrel{\hbox{\tiny law}}{=}$. Two of such symmetries will be repeatedly used. The first is statistical homogeneity, \begin{veqnarray} \label{eq:1.5} u(x,t)\stackrel{\hbox{\tiny law}}{=} u(x+y,t) \quad \hbox{for all } y. \end{veqnarray} (\ref{eq:1.5}) holds if for all measurable, real-valued functional $\Phi(u(\cdot,t))$ we have \begin{veqnarray} \nonumber \< \Phi(u(\cdot,t))\> = \< \Phi(u(\cdot+y,t))\> \quad \hbox{for all } y. \end{veqnarray} Note that from (\ref{eq:1.1}) and the assumptions on~$f$, it follows that (\ref{eq:1.5}) holds for all times if $u_0(x)\stackrel{\hbox{\tiny law}}{=} u_0(x+y)$, where $u_0(\cdot)=u(\cdot,0)$ is the initial velocity field. Also, if a statistical steady state exists and is unique, then it satisfies (\ref{eq:1.5}). The second symmetry will be referred to as statistical parity invariance, and is related to the invariance of (\ref{eq:1.1}) under the transformation $x\to-x$, $u\to-u$. This implies that \begin{veqnarray} \label{eq:1.6} u(x,t)\stackrel{\hbox{\tiny law}}{=} -u(-x,t), \end{veqnarray} or \begin{veqnarray} \nonumber \< \Phi(u(\cdot,t))\> = \< \Phi(-u(-(\cdot),t))\>. \end{veqnarray} (\ref{eq:1.6}) holds for all times if $u_0(x)\stackrel{\hbox{\tiny law}}{=} -u_0(-x)$. (\ref{eq:1.6}) is also satisfied at statistical steady state. Even though some of our results are formulated in terms of lemmas and theorems, the emphasize here is on the ability to calculate things rather than rigor. We will assume that the initial distribution $Z_0(u_1,x_1,\ldots,u_n,x_n)= Z(u_1,x_1,\ldots,u_n,x_n,0)$ is concentrated on $D(I)$, the space of functions defined on $I\subset \hbox {\Bbb R}} \def\N{\hbox {\Bbb N}$ admitting only jumps discontinuities: \begin{veqnarray} \nonumber D(I)=\{u(x,t), u(x{\scr\pm},t) \,\mbox{exists for all}\, x, \hbox{ with } u(x{\scr+},t)\le u(x{\scr-},t)\}. \end{veqnarray} Note that the existence of stationary states is only established when $I$ is a finite interval on $\hbox {\Bbb R}} \def\N{\hbox {\Bbb N}$ \cite{ekhma99}. We will also assume that the initial velocity field $u_0(\cdot)=u(\cdot,0)$ is independent of $\nu$ and statistically independent of $f(\cdot)$. Finally, unspecified integration ranges are meant to be $(-\infty,+\infty)$. \section{Velocity and velocity differences} \label{sec:3} Let $u^\nu(x,t,f,u_0)$ be the solution of (\ref{eq:1.1}) with forcing $f$, initial data $u_0$. It follows from standard results (for the deterministic case, see \cite{lax57,lax73}) that for fixed~$t$ \begin{veqnarray} \nonumber u^\nu(\cdot,t,f,u_0)\to u(\cdot,t,f,u_0) \quad \hbox{as } \nu\to0, \end{veqnarray} $P\times P_0$-almost surely, in $L^2(I)$, where $I$ is the interval on which (\ref{eq:1.1}) is studied. Hence \begin{veqnarray} \nonumber Z^\nu \to Z, \end{veqnarray} weakly. For the statistically stationary states, it was established in \cite{ekhma99} that, weakly, we have \begin{veqnarray} \nonumber Z_\infty^\nu \to Z_\infty. \end{veqnarray} Therefore to study the inviscid limit of statistical quantities of $u$, such as the PDFs of $u$ and $\delta u(x,y,t)=u(x+y,t)-u(y,t)$, it is enough to consider (\ref{eq:1.2}). It is useful, however, to write this equation in a modified form more convenient for calculations. To this end, let \begin{veqnarray} \nonumber u_\pm(x,t)=u(x\pm,t). \end{veqnarray} For any function $g(u)=g(u(x,t))$, we define two average functions \begin{veqnarray} \label{eq:3.1} {\hbox{$[$}}} \def\br{{\hbox{$]$}} g(u)\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} &=& {\t \frac{1}{2}} (g(u_+)+g(u_-))\\ {\hbox{$[$}}} \def\br{{\hbox{$]$}} g(u)\br_\B &=& \int_0^1\!\! d\beta\ g(u_-+\beta(u_+-u_-)). \end{veqnarray} It is shown in \cite{vol67} that (\ref{eq:1.2}) is equivalent to \begin{veqnarray} \label{eq:3.3} u_t+{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x = f, \end{veqnarray} or, in the language of stochastic differential equations, \begin{veqnarray} \label{eq:3.3b} du=-{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_xdt +dW(x,t). \end{veqnarray} (\ref{eq:3.1}) assigns an unambiguous meaning to the quantity ${\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x$ when the solutions of (\ref{eq:3.3}) develop shocks. Moreover the following chain and product rules hold \begin{veqnarray} \label{eq:3.4} g_x(u)={\hbox{$[$}}} \def\br{{\hbox{$]$}} g_u(u)\br_\B u_x, \end{veqnarray} \begin{veqnarray} \label{eq:3.5} (g(u)h(u))_x={\hbox{$[$}}} \def\br{{\hbox{$]$}} g(u)\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} h_x(u) + {\hbox{$[$}}} \def\br{{\hbox{$]$}} h(u)\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} g_x(u). \end{veqnarray} Similar rules apply for derivatives in $t$. As an example of these rules, we have that the integral of the term ${\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x $ across a shock located at $x=y$ is given by (using ${\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}}={\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_\B$) \begin{veqnarray} \nonumber \int_{y-}^{y+}\!\!dx \ {\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x = \int_{y-}^{y+}\!\!dx \ {\t \frac{1}{2}}(u^2)_x = {\t \frac{1}{2}} (u_+^2-u_-^2). \end{veqnarray} It is also convenient to define \begin{veqnarray} \nonumber \bar u(x,t)&=&{\t \frac{1}{2}} (u_+(x,t)+u_-(x,t)), \\ \nonumber s(x,t)&=&u_+(x,t)-u_-(x,t). \end{veqnarray} In terms of $(\bar u,s)$, the two averages in (\ref{eq:3.1}) are \begin{veqnarray} \nonumber {\hbox{$[$}}} \def\br{{\hbox{$]$}} g(u)\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} &=& {\t \frac{1}{2}} (g(\bar u+s/2)+g(\bar u-s/2))\\ \nonumber {\hbox{$[$}}} \def\br{{\hbox{$]$}} g(u)\br_\B &=& \int_{-1/2}^{1/2} \!\!d\beta\ g(\bar u +\beta s). \end{veqnarray} We will denote by $\{y_j\}$ the set of shock positions at time $t$. Hence the set \begin{veqnarray} \nonumber \{(y_j,\bar u(y_j,t),s(y_j,t))\}, \end{veqnarray} quantifies the shocks at time~$t$. \subsection{Master equation for the inviscid Burgers equation} \label{sec:3.1} We now turn to $R(u,x,t)$, the one-point PDF of the solution of (\ref{eq:3.3}). We have \begin{theorem} \label{th:3.1} $R$ satisfies \begin{veqnarray} \label{eq:3.6} R_t=-u R_x -\int du'\ K(u-u') R_x(u',x,t)+ B_0 R_{uu}+G, \end{veqnarray} where $B_0=B(0)$, $K(u)=(H(u)-H(-u))/2$, $H(\cdot)$ is the Heaviside function and $G(u,x,t)$ is given by \begin{veqnarray} \label{eq:3.7} G(u,x,t)=\left( \int_{-\infty}^0\!\! d s \int_{u+s/2}^{u-s/2}\!\! d\bar u \ (u-\bar u) \varrho(\bar u,s,x,t)\right)_u. \end{veqnarray} Here $\varrho(\bar u,s,x,t)$ is defined such that $\varrho(\bar u,s,x,t)d \bar u ds dx$ gives the average number of shocks in $[x,x+dx)$ with $ \bar u(y,t) \in [\bar u, \bar u+d\bar u)$ and $s(y,t) \in [s, s+ds)$, where $y\in[x,x+dx)$ is the shock location. \end{theorem} \demo{Remark 1} $G$ can be referred to as a {\em dissipative anomaly}. It can be written more explicitly as \begin{veqnarray} \label{eq:3.8} G(u,x,t)&=&\frac{1}{2} \int_{-\infty}^0\!\! d s \ s \bigl(\varrho(u-s/2,s,x,t)+\varrho(u+s/2,s,x,t)\bigr)\\ &-&\int_{-1/2}^{1/2}\!\!d\beta\int_{-\infty}^0\!\! ds \ s \varrho(u+\beta s,s,x,t). \end{veqnarray} \enddemo \demo{Remark 2} $\varrho (\bar u,s,x,t)$ can be equivalently defined as \begin{veqnarray} \nonumber \varrho (\bar u,s,x,t)= \int \frac{d\lambda d\mu}{(2\pi)^2}\ {\rm e}^{i\lambda \bar u+i\mu s} \,\hat \varrho(\lambda,\mu,x,t). \end{veqnarray} where \begin{veqnarray} \nonumber \hat \varrho (\lambda,\mu,x,t)= \biggl\< \sum_j{\rm e}^{-i\lambda \bar u(y_j,t)-i\mu s(y_j,t)} \delta(x-y_j)\biggr\>. \end{veqnarray} \enddemo ~ (\ref{eq:3.6}), (\ref{eq:3.7}) do not require statistical homogeneity. For homogeneous situations, these equations simplify. Since $R_x=0$ (\ref{eq:3.6}) reduces to \begin{veqnarray} \label{eq:3.6.b} R_t=B_0 R_{uu}+G, \end{veqnarray} where $G(u,x,t)=G(u,t)$. Since the shock characteristics are independent of its location, we have \begin{veqnarray} \nonumber \varrho(\bar u,s,x,t)=\rho S(\bar u,s,t) \end{veqnarray} where $\rho=\rho(t) $ is the number density of shocks, $S(\bar u,s,t)$ is the PDF of $(\bar u(y_0,t),s(y_0,t))$, conditional on the property that $y_0$ is a shock position (because of the statistical homogeneity, $y_0$ is a dummy variable). Thus, (\ref{eq:3.7}) reduces to \begin{veqnarray} \label{eq:3.10} G(u,t)=\rho\left( \int_{-\infty}^0\!\! d s \int_{u+s/2}^{u-s/2}\!\! d\bar u \ (u-\bar u) S(\bar u,s,t)\right)_u. \end{veqnarray} At statistical stationary state, $R_t=0$ in (\ref{eq:3.6.b}), and one readily verifies from this equation that \begin{veqnarray} \label{3.12.b} R_\infty(u)=\frac{\rho}{2B_0} \int_{-\infty}^0\!\! d s \int_{u+s/2}^{u-s/2}\!\! d\bar u \ \left(\frac{s^2}{4}-(u-\bar u)^2\right) S_\infty(\bar u,s), \end{veqnarray} Clearly, $R_\infty\ge 0$ since $S_\infty(\bar u,s)\ge0$. In addition, by direct computation we obtain \begin{veqnarray} \nonumber \int du \ R_\infty(u)=-\frac{\rho}{12B_0}\< s^3\>. \end{veqnarray} Thus, from the requirement that $R_\infty$ be normalized to unity, we get \begin{veqnarray} \label{eq:3.15b} B_0=-\frac{\rho}{12} \< s^3\>. \end{veqnarray} \begin{proof}[Proof of Theorem \ref{th:3.1}] Let $\theta(\lambda,x,t) = {\rm e}^{-i\lambda u(x,t)}$. $\< \theta\>$ is the characteristic function of $u(x,t)$ and $R$ is given by \begin{veqnarray} \nonumber R(u,x,t)=\int \frac{d\lambda}{2\pi} {\rm e}^{i\lambda u} \<\theta(\lambda,x,t)\>. \end{veqnarray} Using Ito calculus, \begin{veqnarray} \nonumber dW(x,t)dW(y,t)=2B(x-y)dt, \end{veqnarray} it follows from (\ref{eq:3.3b}), (\ref{eq:3.4}), (\ref{eq:3.5}) that \begin{veqnarray} \nonumber d\theta = (-i\lambda{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x{\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B -\lambda^2 B_0\theta)dt -i\lambda \theta dW(x,t). \end{veqnarray} Thus \begin{veqnarray} \nonumber \<\theta\>_t = i\lambda \<{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x{\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B\> -\lambda^2 B_0 \<\theta\>. \end{veqnarray} Note that the terms involving the force contain $\theta$, not ${\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B$. This is because $\theta={\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B$ except when there is a shock, and this is a set of zero probability. Of course this argument does not apply for the convection term since $u_x$ is infinite at shocks. To average the convective term $i\lambda {\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x{\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B$ we use \begin{veqnarray} \nonumber \theta_x= -i\lambda u_x {\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B \end{veqnarray} to get \begin{veqnarray} \nonumber i\lambda \bigl\<{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x {\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B\bigr\> = - \bigl\<{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} \theta_x\bigr\> = - \bigl\< u \theta\bigr\>_x + \bigl\< u_x {\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}}\bigr\>. \end{veqnarray} For convenience we write this equation as \begin{veqnarray} \nonumber i\lambda \bigl\<{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} u_x {\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B\bigr\> =-i\<\theta\>_{x\lambda} +i\lambda^{-1}\<\theta\>_x +\bigl\< u_x \bigl({\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}}-{\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B \bigr)\bigr\>. \end{veqnarray} Combining these expressions gives: \begin{veqnarray} \label{eq:3.11.0} \<\theta\>_t =-i\<\theta\>_{x\lambda} +i\lambda^{-1}\<\theta\>_x-\lambda^2 B_0 \<\theta\> +\hat G, \end{veqnarray} where $\hat G(\lambda,x,t)$ is given by \begin{veqnarray} \nonumber \hat G(\lambda,x,t)=\bigl\< u_x \bigl({\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}}-{\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B \bigr) \bigr\>. \end{veqnarray} To proceed with the evaluation of $\hat G$, note that the only contributions to this term are from the shocks, since ${\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}}={\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta\br_\B$ except at shocks. Let $\{y_j\}$ denote the positions of the shocks at time $t$. Since $u_x(y_j,t)=s(y_j,t)\delta(x-y_j)$ at the shocks, $\hat G$ can be understood as \begin{veqnarray} \label{eq:3.11} &&\hat G(\lambda,x,t)\\ &=&\biggl\< \sum_j s_j \delta(x-y_j) \Bigl({\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta(\lambda,y_j,t)\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} -{\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta(\lambda,y_j,t)\br_\B \Bigr) \biggr\>, \end{veqnarray} where $s_j=s(y_j,t)$. Using (\ref{eq:3.1}), \begin{veqnarray} \nonumber &&\hat G(\lambda,x,t)\\ &=&\biggl\< \sum_j s_j \delta(x-y_j) {\rm e}^{-i\lambda \bar u_j} \biggl({\rm e}^{i\lambda s_j/2}+{\rm e}^{-i\lambda s_j/2} -2\int_{-1/2}^{1/2}\!\! d\beta\ {\rm e}^{-i\lambda\beta s_j}\biggr)\biggr\>, \end{veqnarray} where $\bar u_j=\bar u(y_j,t)$. Using $\varrho(\bar u,s,x,t)$ this average is \begin{veqnarray} \nonumber &&\hat G(\lambda,x,t)\\ &=&\int d\bar u \int_{-\infty}^{0}\!\!ds \ \frac{s}{2}{\rm e}^{-i\lambda \bar u} \biggl({\rm e}^{i\lambda s/2}+{\rm e}^{-i\lambda s/2} -2\int_{-1/2}^{1/2}\!\! d\beta\ {\rm e}^{-i\lambda \beta s}\biggr) \varrho(\bar u,s,x,t), \end{veqnarray} Going back to the variable $u$, we get (\ref{eq:3.8}), hence (\ref{eq:3.6}). \end{proof} \demo{Remark} Under the assumption of ergodicity with respect to spatial translations, an alternative derivation of (\ref{eq:3.10}) is to go back to (\ref{eq:3.11}) and use the equivalence between ensemble average and spatial average. Then \begin{veqnarray} \nonumber &&\hat G(\lambda,t)\\ &=&\lim_{L\to\infty}\frac{1}{2L}\int_{-L}^L\!\!dx\ \sum_j s_j \delta(x-y_j) \Bigl({\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta(\lambda,y_j,t)\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} -{\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta(\lambda,y_j,t)\br_\B \Bigr) \biggr\>\\ &=&\lim_{L\to\infty}\frac{N}{2L}\frac{1}{N} \sum_{j=1}^N s_j \Bigl({\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta(\lambda,y_j,t)\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}}- {\hbox{$[$}}} \def\br{{\hbox{$]$}}\theta(\lambda,y_j,t)\br_\B \Bigr), \end{veqnarray} where $N$ is the number of shocks in the interval $[-L,L]$. Using the ergodicity again it follows that the sum is equal to \begin{veqnarray} \nonumber &&\hat G(\lambda,t)\\ &=&\rho \int d\bar u \int_{-\infty}^{0}\!\!ds \frac{s}{2}{\rm e}^{-i\lambda \bar u} \biggl({\rm e}^{i\lambda s/2}+{\rm e}^{-i\lambda s/2} -2\int_{-1/2}^{1/2}\!\! d\beta\ {\rm e}^{-i\lambda\beta s}\biggr) \,S(\bar u,s,t), \end{veqnarray} where we used $\lim_{L\to\infty} N/(2L)=\rho $. Going back to the variable $u$, we get (\ref{eq:3.10}). \enddemo \subsection{An alternative derivation} \label{sec:3.2} The derivation of (\ref{eq:3.6}) (or (\ref{eq:3.6.b}) for statistically homogeneous situations) given above is rigorous, but rather unintuitive. In fact, the effect of the viscous term is buried in the definition of the two averages in (\ref{eq:3.1}), and it is not clear at all at this stage whether (\ref{eq:3.6}) arises in the limit of the equation for $R^\nu$ as $\nu\to0$. Assuming statistical homogeneity, recall that $R^\nu$ satisfies (see (\ref{eq:A.5}) in the Appendix) \begin{veqnarray} \nonumber R^\nu_t=B_0 R^\nu_{uu} -\nu \bigl(\< u_{xx}|u\> R^\nu\bigr)_{u}. \end{veqnarray} Here we give another, less rigorous but more intuitive, derivation of (\ref{eq:3.6.b}) and (\ref{eq:3.10}) by working with the equation for $R^\nu$, and calculating directly the limit of the viscous term as $\nu\to0$. In particular, we will compute explicitly that the dissipative anomaly (\ref{eq:3.10}) is given by \begin{veqnarray} \label{eq:3.9} G(u,t) = -\lim_{\nu\to0} \nu \bigl(\< u_{xx} |u\> R\bigr)_u. \end{veqnarray} This will give strong support to the claim that $\underline R=\lim_{\nu\to0}R^\nu$ satisfies (\ref{eq:3.6.b}). Assuming spatial ergodicity, the average of the dissipative term can be expressed as \begin{veqnarray} \label{eq:3.12} \nu \< u_{xx}|u\> R^\nu &=&\nu \< u_{xx}(x,t) \delta(u-u(x,t))\>\\ &=&\nu\lim_{L\to\infty}\frac{1}{2L} \int_{-L}^{L} dx\ u_{xx}(x,t) \delta(u-u(x,t)). \end{veqnarray} Clearly, in the limit as $\nu\to0$ only small intervals around the shocks will contribute to the integral. In these intervals, boundary layer analysis can be used to obtain an accurate approximation of $u(x,t)$. The basic idea is to split $u$ into the sum of an inner solution near the shock and an outer solution away from the shock, and using systematic matched asymptotics to construct uniform approximation of $u$ (for details see, e.g., \cite{goxi92}). For the outer solution, we look for an approximation in the form of a series in $\nu$: \begin{veqnarray} \nonumber u=u^{\hbox{\tiny out}}} \def\inn{{\hbox{\tiny in}}= u_0+\nu u_1+O(\nu^2). \end{veqnarray} Then $u_0$ satisfies \begin{veqnarray} \nonumber {u_0}_t+u_0 {u_0}_x=f, \end{veqnarray} i.e. Burgers equation without the dissipation term. In order to deal with the inner solution around the shock, let $y=y(t)$ be the position of a shock, and define the stretched variable $z=(x-y)/\nu$ and let \begin{veqnarray} \nonumber u^\inn(x,t)=v\left(\frac{x-y}{\nu}+\delta,t\right), \end{veqnarray} where $\delta$ is a perturbation of the shock position to be determined later. Then, $v$ satisfies \begin{veqnarray} \label{eq:3.13} \nu v_t +(v-\bar u+\nu \gamma)v_z =v_{zz}+\nu f, \end{veqnarray} where $\bar u =dy/dt$, $\gamma = d\delta/dt$ and, to $O(\nu^2)$, $\nu f$ can be evaluated at $x=y$ (and can thus be considered as a function of $t$ only). We study (\ref{eq:3.13}) by regular perturbation analysis. We look for a solution in the form \begin{veqnarray} \nonumber v= v_0 +\nu v_1+O(\nu^2). \end{veqnarray} To leading order, from (\ref{eq:3.13}) we get for $v_0$ the equation \begin{veqnarray} \nonumber (v_0-\bar u){v_0}_z ={v_0}_{zz}. \end{veqnarray} The boundary condition for this equation arises from the matching condition with $u^{\hbox{\tiny out}}} \def\inn{{\hbox{\tiny in}}=u_0+\nu u_1 +O(\nu^2)$: \begin{veqnarray} \nonumber \lim_{z\rightarrow \pm\infty} v_0 = \lim_{x\to y} u_0 \equiv \bar u\pm\frac{s}{2}, \end{veqnarray} where $s=s(t)$ is the shock strength. It is understood that for small $\nu$ matching takes place for small values of $|x-y|$ and large values of $|z|=|x-y|/\nu$. This gives \begin{veqnarray} \nonumber v_0= \bar u-\frac{s}{2} \tanh \left(\frac{s z}{4}\right). \end{veqnarray} These results show that, to $O(\nu)$, (\ref{eq:3.12}) can be estimated as \begin{veqnarray} \nonumber \nu \< u_{xx}|u\> R^\nu =\nu\lim_{L\to\infty}\frac{N}{2L} \frac{1}{N} \sum_{j} \int_{\Omega_j}\!\! dx\ u^\inn_{xx}\, \delta(u-u^\inn(x,t)), \end{veqnarray} where $\Omega_j$ is a layer centered at $y_j$ with width $\gg O(\nu)$. Going to the stretched variable $z=(x-y)/\nu$, and taking the limit as $L\to\infty$, we get \begin{veqnarray} \nonumber \nu \< u_{xx}|u\> R^\nu =\rho \int d\bar u \int_{-\infty}^{0}\!\!ds \ S(\bar u,s;t)\int_{-\infty}^{+\infty}\!\! dz \ {v_0}_{zz}(z,t) \delta(u-v_0(z,t)), \end{veqnarray} where $S(\bar u,s;t)$ is the PDF of $(\bar u(y_0,t),s(y_0,t))$ conditional on $y_0$ being a shock location. The $z$-integral can be evaluated exactly using \begin{veqnarray} \nonumber dz {v_0}_{zz}=dv_0\frac{{v_0}_{zz}}{{v_0}_z}= dv_0 (v_0-\bar u), \end{veqnarray} where we used the equation $(v_0-\bar u){v_0}_z ={v_0}_{zz}$. This leads to \begin{veqnarray} \nonumber \nu \< u_{xx} |u\> R^\nu =-\rho \int d\bar u \int_{-\infty}^{0}\!\!ds \ S(\bar u,s;t)\int_{\bar u+s/2}^{\bar u-s/2} \!\! dv_0\ (v_0-\bar u) \delta(u-v_0). \end{veqnarray} Hence, \begin{veqnarray} \nonumber \lim_{\nu\to0}\nu \< u_{xx}|u\> R^\nu= -\rho\int_{-\infty}^0\!\! d s \int_{u+s/2}^{u-s/2}\!\! d\bar u \ (u-\bar u) S(\bar u,s;t). \end{veqnarray} Using this expression in (\ref{eq:3.9}), we get (\ref{eq:3.7}). \subsection{Computing the anomalies} \label{sec:3.3} Here we derive equations for the moments of $R$. For simplicity, consider a statistically homogeneous case, and assume that $u(x,t)\stackrel{\hbox{\tiny law}}{=} -u(-x,t)$. Then $S(\bar u,s,t)=S(-\bar u,s,t)$, $G(u,t)=G(-u,t)$, $R(u,t)=R(-u,t)$. This means that all moments of odd orders of $R$ are zero. For the moments of even orders, we get from (\ref{eq:3.6.b}) the following equations ($n\in\N_0$): \begin{veqnarray} \nonumber \frac{d}{dt}\< u^{2n}\> = 2n(2n-1)B_0\< u^{2n-2}\>+h_{2n}, \end{veqnarray} where $h_{2n}$ is the anomaly term \begin{veqnarray} \nonumber h_{2n} &=& \int du \ u^{2n} G \\ &=&-\frac{\rho}{2^{2n}(2n+1)} \Bigl(\< (2\bar u-s)^{2n}(\bar u +ns)\>- \< (2\bar u+s)^{2n}(\bar u -ns)\>\Bigr). \end{veqnarray} ($h_{2n+1}=0$ by parity). An alternative definition of $h_{2n}$ is \begin{veqnarray} \nonumber h_{2n}= 2n \lim_{\nu\to0}\nu \< u^{2n-1} u_{xx}\> = - 2n(2n-1) \lim_{\nu\to0}\nu \< u^{2n-2} u^2_x\>. \end{veqnarray} This gives for instance \begin{veqnarray} \nonumber h_2=-2\lim_{\nu\to0}\nu \< u_x^2\> =\frac{\rho}{6} \< s^3\>. \end{veqnarray} At statistical steady state, it gives \begin{veqnarray} \nonumber \< u^{2n}\>=\frac{C_{2n}\rho}{B_0} \Bigl(\< (2\bar u-s)^{2n+2}(\bar u +(n+1)s)\>- \< (2\bar u+s)^{2n+2}(\bar u -(n+1)s)\>\Bigr), \end{veqnarray} where $C_n=n!/2^{n+2}(n+3)!$. These expressions can also be obtained from (\ref{3.12.b}). In particular, for $n=2$, we obtain again (\ref{eq:3.15b}). \subsection{Multi-point PDF} \label{sec:3.5} We now turn to $Z(u_1,x_1,\ldots,u_n,x_n,t)$, the multi-point PDF of $u$. We have \begin{theorem} \label{th:4} $Z$ satisfies \begin{veqnarray} \label{eq:3.150} Z_t&=&-\sum_{p=1}^n u_pZ_{x_p}+\sum_{p,q=1}^n B(x_p-x_q)Z_{u_pu_q}\\ &&-\sum_{p=1}^n\int du'\ K(u_p-u') Z_{x_p}(u_1,x_1,\ldots,u',x_p,\ldots,u_n,x_n,t)\\ &&+\sum_{p=1}^n G(u_p,x_p,u_2,x_2\ldots,u_{p-1},x_{p-1},u_{p+1},x_{p+1}, \ldots,u_n,x_n,t), \end{veqnarray} where $K(u)=(H(u)-H(-u))/2$, $H(\cdot)$ is the Heaviside function and $G$ is given by \begin{veqnarray} \label{eq:3.16} &&G(u_1,x_1,\ldots,u_n,x_n,t)\\ &&=\biggl( \int_{-\infty}^0\!\! d s \int_{u_1+s/2}^{u_1-s/2}\!\! d\bar u \ (u_1-\bar u)\\[-6pt] &&\qquad\qquad\qquad\times \varrho(\bar u,s,x_1,u_2,x_2,\ldots,u_n,x_n,t)\biggr)_{u_1}. \end{veqnarray} Here $\varrho(\bar u,s,x_1,u_2,x_2\ldots,u_n,x_n,t)$ is defined such that \begin{veqnarray} \nonumber \varrho(\bar u,s,x_1,u_2,x_2,\ldots,u_n,x_nt)d \bar u dsdu_2\cdots du_n dx, \end{veqnarray} gives the average number of shocks in $[x_1,x_1+dx)$ with $ \bar u(y,t) \in [\bar u, \bar u+d\bar u)$, $s(y,t) \in [s, s+ds)$ where $y\in[x_1,x_1+dx)$ is a shock location, and with $u(x_2,t)\in [u_2, u_2+du_2)$,\ldots, $u(x_n,t)\in [u_n, u_n+du_n)$. \end{theorem} We will omit the proof of Theorem~\ref{th:4} since it is a straightforward generalization of the one given for Theorem~\ref{th:3.1}. ~ For the two-point PDF, $Z(u_1,x_1,u_2,x_2,t)$, (\ref{eq:3.150}) becomes \begin{veqnarray} \label{eq:3.15} Z_t&=&-u_1Z_{x_1} -\int du'\ K(u_1-u') Z_{x_1}(u',x_1,u_2,x_2,t)\\ && -u_2Z_{x_2} - \int du'\ K(u_2-u') Z_{x_2}(u_1,x_1,u',x_2,t)\\ &&+B_0 Z_{u_1u_1}+B_0 Z_{u_2u_2}+2B(x_1-x_2)Z_{u_1u_2}\\ &&+G(u_1,x_1,u_2,x_2,t)+G(u_2,x_2,u_1,x_1,t). \end{veqnarray} Assuming statistical homogeneity, (\ref{eq:3.15}) can be simplified. First, \begin{veqnarray} \nonumber Z(u_1,y,u_2,x+y,t)=Z(u_1,0,u_2,x,t)\equiv Z(u_1,u_2,x,t). \end{veqnarray} Secondly, \begin{veqnarray} \nonumber \varrho(\bar u,s,y,u,x+y,t)=\rho T(\bar u,s,u,x,t), \end{veqnarray} where $\rho $ is the number density of shocks, and $T(\bar u,s,u,x,t)$ is the PDF of \begin{veqnarray} \nonumber (\bar u(y_0,t),s(y_0,t),u(y_0+x,t)), \end{veqnarray} conditional on $y_0$ being a shock position. Thus for statistically homogeneous situations, (\ref{eq:3.15}) reduces to the following equation for $Z(u_1,u_2,x,t)$ \begin{veqnarray} \label{eq:3.18} Z_t&=&-(u_2-u_1)Z_{x}\\ &&- \int du'\ K(u_2-u')Z_x(u_1,u',x,t)\\ &&+\int du'\ K(u_1-u')Z_x(u',u_2,x,t)\\ &&+B_0 Z_{u_1u_1}+B_0 Z_{u_2u_2}+2B(x)Z_{u_1u_2}\\ &&+G(u_1,u_2,x,t)+G(u_2,u_1,-x,t), \end{veqnarray} where \begin{veqnarray} \label{eq:3.19} &&G(u_1,u_2,x,t)\\ &=&\rho \biggl( \int_{-\infty}^0\!\! d s \ s \int_{u_1+s/2}^{u_1-s/2}\!\! d\bar u \ (u_1-\bar u)T(\bar u,s,u_2,x,t) \biggr)_{u_1}. \end{veqnarray} \subsection{Velocity difference and structure functions} \label{sec:3.6} Assume statistical homogeneity, and let $Z^\delta(\delta u,x,t)$ be the PDF of the velocity difference $\delta u(x,y,t) =u(x+y,t)-u(y,t)$. $Z^\delta(w,x,t)$ is related to $Z(u_1,u_2,x,t)$ by \begin{veqnarray} \nonumber Z^\delta(w,x,t) =\int du\ Z\left(u-\frac{w}{2},u+\frac{w}{2},x,t\right). \end{veqnarray} It then follows immediately from (\ref{eq:3.18}) and (\ref{eq:3.19}) that \begin{corollary} \label{th:3.4} $Z^\delta$ satisfies \begin{veqnarray} \label{eq:3.20} Z^\delta_t&=&-wZ^\delta_{x}-2\int dw'\ H(w'-w) Z^\delta_x(w',x,t)\\ &&+2(B_0-B(x)) Z^\delta_{ww}+G^\delta(w,x,t), \end{veqnarray} where $K(w)=(H(w)-H(-w))/2$, $H(\cdot)$ is the Heaviside function and \begin{veqnarray} \label{eq:3.21} G^\delta(w,x,t)&=&\int du\ G\left(u-\frac{w}{2},u+\frac{w}{2},x,t\right)\\ &+&\int du\ G\left(u+\frac{w}{2},u-\frac{w}{2},-x,t\right). \end{veqnarray} \end{corollary} \demo{Remark} $G^\delta$ can be put into a form which is more convenient for the calculations. Let \begin{veqnarray} \nonumber \delta u_+(x,y_0,t)=u(y_0+|x|,t)-u_+(y_0,t),\\ \delta u_-(x,y_0,t)= u_-(y_0,t)-u(y_0-|x|,t), \end{veqnarray} and let $U_\pm(s,\delta u_\pm,x,t)$ be the PDFs of $(s(y_0,t),\delta u_\pm(x,y_0,t))$ conditional on $y$ being a shock position. Then, $G^\delta$ can be expressed as \begin{veqnarray} \nonumber G^\delta(w,x,t)=G^\delta_+(w,x,t)+G^\delta_-(w,x,t), \end{veqnarray} where \begin{veqnarray} \nonumber G^\delta_\pm(w,x,t)&=&\frac{\rho }{2}\int_{-\infty}^0\!\! ds \ s\bigl( U_\pm(s,{\rm sgn}(x) w-s,x,t)+U_\pm(s,{\rm sgn}(x) w,x,t)\bigr)\\ &-&\rho \int_{-\infty}^0\!\! ds \ s\int_0^1d\beta \ U_\pm (s,{\rm sgn}(x) w-\beta s,x,t). \end{veqnarray} \enddemo ~ We now consider some consequences of (\ref{eq:3.20}). We have the following two theorems: \begin{theorem} \label{th:3.5} In the limit as $x\to0$, \begin{veqnarray} \label{eq:3.22} &Z^\delta(w,x,t)=\delta(w)-|x|(\rho\delta(w)+\< \xi\> \delta^1(w)-\rho S(w,t))+o(x),\\[4pt] &{\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle Z^\delta(w,x,t)= (1-\rho |x|)\frac{1}{x} Q\left(\frac{w}{x},t\right) +|x|\rho S(w,t)+o(x),} \end{veqnarray} where $\delta^1(w)=d\delta(w)/dw$, $o(x)$ must be interpreted in the sense of weak convergence. Here $Q(\xi,t)$ is the PDF of $\xi(x,t)$, the regular part of the velocity gradient, i.e. \begin{veqnarray} \nonumber u_x(x,t)= \xi(x,t)+\sum_j s(y_j,t) \delta (y-y_j), \end{veqnarray} and $S(s,t)$ is the PDF of $s(y_0,t)$ conditional on $y_0$ being a shock location. \end{theorem} \begin{theorem}[Structure function scaling] \label{th:3.6} Let \begin{veqnarray} \label{eq:3.aa} \< |\delta u|^a\>= \int dw \ |w|^a Z^\delta(w,x,t). \end{veqnarray} In the limit as $x\to0$, \begin{veqnarray} \label{eq:3.aaa} \< |\delta u|^a\> = \left\{ \begin{array}{ll} {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle |x|^{a} \< |\xi|^a\>+o(x^a)\quad}& \mbox{if}\ \ 0\le a<1, \\[6pt] {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle |x| \bigl( \< |\xi|\>+\rho \< |s|\>\bigr) + o(x)\quad}&\mbox{if}\ \ a=1,\\[6pt] {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle |x| \rho \< |s|^a\> + o(x)\quad}&\mbox{if}\ \ 1 < a. \end{array}\right. \end{veqnarray} \end{theorem} In terms of the moments, (\ref{eq:3.aaa}) is ($n\in\N_0$) \begin{veqnarray} \label{eq:3.bbb} \< \delta u^{2n}\>&=&|x| \rho \< s^{2n}\> + o(x),\\ \< \delta u^{2n+1}\>&=&x \rho \< s^{2n+1}\> + o(x). \end{veqnarray} The first moment satisfies $\< \delta u\> =0$ for all $(x,t)$. This is a consequence of statistical homogeneity and it is readily verified since multiplying (\ref{eq:3.20}) by $w$ and integrating gives in $\< \delta u\>_t=0$. An equation for $Q(\xi,t)$ will be derived in Sec.~\ref{sec:4}. To interpret (\ref{eq:3.22}) note that $Z^\delta$ can be decomposed into \begin{veqnarray} \nonumber Z^\delta(w,x,t)&=& p_{ns}(x,t) Z^\delta(w,x,t|\hbox{no shock})\\ &+&(1-p_{ns}(x,t)) Z^\delta(w,x,t|\hbox{shock}), \end{veqnarray} where $p_{ns}(x,t)$ is the probability that there is no shock in $[y,y+x)$, $Z^\delta(w,x,t|\hbox{no shock})$ is the PDF of $\delta u(x,y,t)$ conditional on the property that there is no shock in $[y,y+x)$, $Z^\delta(w,x,t|\hbox{shock})$ is the PDF of $\delta u(x,y,t)$ conditional on the property that there is at least one shock in $[y,y+x)$. Since, by definition of~$\rho$ we have \begin{veqnarray} \nonumber p_{ns} = 1 -\rho |x| + o(x), \end{veqnarray} (\ref{eq:3.22}) states that \begin{veqnarray} \nonumber Z^\delta(w,x,t|\hbox{no shock}) &=& (1-\rho |x|)\frac{1}{x} Q\left(\frac{w}{x},t\right)+o(x),\\ Z^\delta(w,x,t|\hbox{shock})&=&S(w,t)+o(1). \end{veqnarray} This is consistent with the picture that $\delta u(x,y,t)=x \xi(y,t)+o(x)$ if there is no shock in $[y,y+x)$, and $w(x,y,t)=s(y_0,t)+o(1)$ if $y_0\in[y,x+y)$ is a shock position. \begin{proof}[Proof of Theorem \ref{th:3.5}] We will consider the case $x>0$. The case $x<0$ can be treated similarly. Note first that, in the limit as $x\to0$, we have \begin{veqnarray} \nonumber x Z^\delta(x\xi,x,t)\to Q(\xi,t), \end{veqnarray} weakly. We postpone the proof of this fact until Sec.~\ref{sec:4}. It implies that \begin{veqnarray} \nonumber Z^\delta(w,x,t)= \delta(w)+o(1), \end{veqnarray} weakly. Define \begin{veqnarray} \nonumber A(w,t)= \lim_{x\to0} x^{-1}(Z^\delta(w,x,t)-\delta(w))= \lim_{x\to0} {Z}_x^\delta(w,x,t). \end{veqnarray} Taking the limit as $x\to0$ in the equation for $Z^\delta$, it follows that $A$ satisfies \begin{veqnarray} \label{eq:3.a.1} 0=-wA-2\int dw'\ H(w-w') A(w',x,t)+B(w,t), \end{veqnarray} where we used $\lim_{x\to0}(B_0-B(x))=0$ and we defined \begin{veqnarray} \nonumber B(w,t)=\lim_{x\to0}G^\delta(w,x,t). \end{veqnarray} To evaluate $B$ note that as $x\to0$ \begin{veqnarray} \nonumber \delta u_\pm (x,y_0,t)\to0, \end{veqnarray} almost surely. This implies that, as $x\to0$, \begin{veqnarray} \nonumber U_\pm(s,w,x,t) \to S(s,t)\delta(w), \end{veqnarray} where $S(s,t)$ is the PDF of $s(y_0,t)$ conditional on $y_0$ being a shock location. Hence, from the expression for $G^\delta$, \begin{veqnarray} \nonumber B(w,t)=\rho w S(w,t)+\rho \< s\> \delta(w) +2\rho \int_{-\infty}^w \!\! dw' S(w't)- 2\rho H(w), \end{veqnarray} where $H(\cdot)$ is the Heaviside function and we used $S(s,t)=0$ for $s>0$ since $s(y_0,t)\le0$. Inserting this expression in (\ref{eq:3.a.1}), the solution of this equation is \begin{veqnarray} \nonumber A(w,t)=-\delta(w)+\rho \< s\> \delta^1(w)+\rho S(w,t). \end{veqnarray} Here $\delta^1(w)=d\delta(w)/dw$ and we used the identity $w\delta^1(w)=-\delta(w)$. Using (\ref{eq:4.3.1}), $\rho\< s\>=-\< \xi\>$, this can be restated as \begin{veqnarray} \nonumber A(w,t)=-\delta(w)-\< \xi\> \delta^1(w)+\rho S(w,t). \end{veqnarray} Hence, combining the above results, we have \begin{veqnarray} \nonumber Z^\delta(w,x,t)=\delta(w)-x(\delta(w)+\< \xi\> \delta^1(w)- \rho S(w,t))+o(x), \end{veqnarray} weakly. This establishes the first equation in (\ref{eq:3.22}) for $x>0$. Reorganizing this expression as \begin{veqnarray} Z^\delta(w,x,t) &=&(1-\rho x)(\delta(w)-x\< \xi\> \delta^1(w))+ x\rho S(w,t)+o(x), \end{veqnarray} weakly, and using the identity \begin{veqnarray} \nonumber \delta(w)-x\< \xi\> \delta^1(w)=\frac{1}{x} Q\left(\frac{w}{x},t\right)+o(x), \end{veqnarray} establishes the second equation in (\ref{eq:3.22}) for $x>0$. \end{proof} \begin{proof}[Proof of Theorem \ref{th:3.6}] We will prove (\ref{eq:3.aaa}) directly for moments of integer order higher than one. For other values of $a$ (\ref{eq:3.aaa}) follows from (\ref{eq:3.22}) and the fact that the tails are controlled by higher order moments. Note first that for $a>0$ \begin{veqnarray} \nonumber \lim_{x\to0} \< |\delta u|^a\> =0, \end{veqnarray} since $\delta u(x,y,t)\to0$ almost surely as $x\to0$. Now, multiply (\ref{eq:3.20}) by $w^{n}$ ($n\in \N$, $n\ge 2$), integrate and take the limit as $x\to0\pm$. The result is \begin{veqnarray} \nonumber 0=-a^\pm_{n+1}+\frac{2}{n+1} a^\pm_{n+1}+b^\pm_{n}, \end{veqnarray} where \begin{veqnarray} \nonumber a^\pm_n=\lim_{x\to 0\pm} x^{-1} \< \delta u^n\> =\lim_{x\to 0\pm}\< \delta u^n\>_x, \end{veqnarray} \begin{veqnarray} \nonumber b^\pm_n=\lim_{x\to 0\pm} \int dw \ w^n G^\delta(w,x,t). \end{veqnarray} Note that \begin{veqnarray} \nonumber \int dw \ w^n G^\delta(w,x,t) &=& \frac{\rho}{2} \< s(\delta u_++{\rm sgn}(x) s)^n\> +\frac{\rho}{2} \< s\delta u_+^n\>\\ &-&\rho \int_0^1 \!\! d\beta\ \< s(\delta u_++\beta{\rm sgn}(x) s)^n\>\\ &+&\frac{\rho}{2} \< s( \delta u_-+{\rm sgn}(x)s)^n\> +\frac{\rho}{2} \< s\delta u_-^n\>\\ &-&\rho \int_0^1 \!\! d\beta\ \< s(\delta u_-+\beta{\rm sgn}(x) s)^n\>. \end{veqnarray} Since $\delta u_\pm (x,y_0,t)\to 0$ almost surely as $x\to0$, we have \begin{veqnarray} \nonumber b^\pm_n= (\pm1)^{n} \rho \< s^{n+1}\> \left(1-\frac{2}{n+1}\right). \end{veqnarray} Inserting this expression in the equation for $a^\pm_n$ gives \begin{veqnarray} \nonumber a^\pm_{n+1}=\lim_{x\to 0\pm} x^{-1} \< \delta u^{n+1}\>= (\pm1)^{n}\rho \< s^{n+1}\>. \end{veqnarray} Thus \begin{veqnarray} \nonumber \< \delta u^{2n}\> = |x|\rho \< s^{2n}\>+o(x),\qquad \< \delta u^{2n+1}\> = x\rho \< s^{2n+1}\>+o(x). \end{veqnarray} This proves (\ref{eq:3.bbb}). We now prove (\ref{eq:3.aaa}) for $0\le a\le1$. The proof for other values of $a$ is similar. Let \begin{veqnarray} \nonumber f^\delta(w,x,t)&=&(1-\rho |x|)\frac{1}{x} Q\left(\frac{w}{x},t\right) +|x|\rho S(w,t),\\[4pt] g^\delta(w,x,t)&=&\delta(w)-|x|(\rho \delta(w)+\< \xi\> \delta^1(w)-\rho S(w,t)), \end{veqnarray} and write for $M>0$ \begin{veqnarray} \nonumber &&\int dw \ |w|^a (Z^\delta-f^\delta)\\ &=& \int_{|w|\le M}\!\! dw \ |w|^a (Z^\delta-f^\delta)+ \int_{|w|> M}\!\! dw \ |w|^a (Z^\delta-f^\delta). \end{veqnarray} The first term at the right hand-side is $o(x)$ because of (\ref{eq:3.22}). To estimate the second term, note that for $M$ large enough \begin{veqnarray} \nonumber \int_{|w|> M}\!\! dw \ |w|^a Z^\delta&\le& \int_{|w|> M}\!\! dw \ w^2 Z^\delta\\ &\le& \left|\int dw \ w^2 (Z^\delta-g^\delta)\right|+ \int_{|w|> M}\!\! dw \ w^2 g^\delta\\ &=&o(x) +|x|\rho \int_{|w|> M}\!\! dw \ w^2 S(w,t)\\ &=&o(x)+O(x)\int_{|w|> M}\!\! dw \ w^2 S(w,t). \end{veqnarray} \begin{veqnarray} \nonumber \int_{|w|> M}\!\! dw \ |w|^a f^\delta&=& |x|^a (1-|x|\rho) \int_{|\xi|> M/x}\!\! d\xi \ |\xi|^a Q(\xi,t) \\ &&+ |x|\rho\int_{w> M}\!\! dw \ |w|^a S(w,t)\\ &=&o(x^a)+O(x)\int_{w> M}\!\! dw \ |w|^a S(w,t). \end{veqnarray} Since $M$ can be made arbitrarily large, we get \begin{veqnarray} \nonumber \int dw \ |w|^a (Z^\delta-f^\delta)\le o(x^a)+\delta_M O(x), \end{veqnarray} where $\delta_M\to0$ as $M\to+\infty$. Noting that \begin{veqnarray} \nonumber \int dw \ |w|^a f^\delta= \left\{ \begin{array}{ll} {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle |x|^{a} \< |\xi|^a\>\quad}& \mbox{if}\ \ 0\le a<1, \\[6pt] {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle |x| \bigl( \< |\xi|\>+\rho \< |s|\>\bigr) \quad}&\mbox{if}\ \ a=1, \end{array}\right. \end{veqnarray} we obtain (\ref{eq:3.aaa}) for $0\le a\le 1$. \end{proof} \section{Velocity gradient} \label{sec:4} We now turn to the study of the PDF of the velocity gradient for the solution of (\ref{eq:3.3}). Since the solutions typically contain discontinuities, it is already an issue whether the PDF for the velocity gradient is well-defined. Heuristically, this is so since $u$ only fails to be differentiable at no more than countably many points. We will therefore be concerned with the PDF of the regular part of the gradient. \subsection{Master equation for the PDF of the velocity gradient} \label{sec:4.1} We focus on the statistically homogeneous case and derive an equation for $Q(\xi,t)$, the PDF of $\xi(x,t)$, defined as the regular part of the velocity gradient, i.e. \begin{veqnarray} \nonumber u_x(x,t)=\xi(x,t)+\sum_j s(y_j) \delta(x-y_j), \end{veqnarray} where $\xi(\cdot,t)\in L^1(I)$. We will prove that \begin{theorem} \label{th:5} $Q$ satisfies \begin{veqnarray} \label{eq:4.1} Q_t=\xi Q +\bigl(\xi^2 Q\bigr)_\xi+ B_1Q_{\xi\xi}+F(\xi,t), \end{veqnarray} where $B_1=-B_{xx}(0)$ and \begin{veqnarray} \label{eq:4.2} F(\xi,t)=\rho\int_{-\infty}^0\!\! ds \ s V(s,\xi,t). \end{veqnarray} Here $V(s,\xi,t)=(V_+(s,\xi,t)+V_-(s,\xi,t))/2$, $V_\pm(s,\xi_\pm,t)$ are the PDFs of \begin{veqnarray} \nonumber (s(y_0,t),\xi_\pm(y_0,t)=u_x(y_0\pm,t)), \end{veqnarray} conditional on the property that $y_0$ is a shock position. \end{theorem} The consequences of (\ref{eq:4.1}) will be studied in Sec.~\ref{sec:4.3}. Note that if $u(x,t)\stackrel{\hbox{\tiny law}}{=} -u(-x,t)$, then \begin{veqnarray} \nonumber (\xi_+(x,t),s(x,t))\stackrel{\hbox{\tiny law}}{=}(\xi_-(-x,t),s(-x,t)), \end{veqnarray} and $V_+(s,\xi,t)=V_-(s,\xi,t)=V(s,\xi,t)$. \begin{proof}[Proof of Theorem \ref{th:5}] Let $Q^\delta (\eta,x,t)$ be the PDF of $\eta(x,y,t)=(u(x+y,t)-u(y,t))/x$. $Q^\delta$ is related to $Z$ and $Z^\delta$ by \begin{veqnarray} \nonumber Q^\delta (\eta,x,t)=x\, Z^\delta(x\eta,x,t) = x \int du \ Z\left(u-\frac{x\eta}{2},u+\frac{x\eta}{2},x,t\right). \end{veqnarray} From (\ref{eq:3.20}) it follows that $Q^\delta $ satisfies \begin{veqnarray} \label{eq:4.3} Q^\delta_t&=&\eta Q^\delta+\bigl(\eta^2 Q^\delta \bigr)_\eta+ B_1(x)Q^\delta_{\eta\eta} -x\eta Q^\delta_x\\ && -2x\int d\eta'\ H(\eta-\eta') Q^\delta_x(\eta',x,t) +F^\delta(\eta,x,t), \end{veqnarray} where $B_1(x)=2(B_0-B(x))/x^2$ and \begin{veqnarray} \nonumber F^\delta(\eta,x,t)= F^\delta_1(\eta,x,t)+F^\delta_2(\eta,x,t)+F^\delta_3(\eta,x,t), \end{veqnarray} with \begin{veqnarray} \nonumber F^\delta_1(\eta,x,t)&=& \rho\int_{-\infty}^0\!\! ds \ s V^\delta \left(s,\eta,x,t\right),\\ F^\delta_2(\eta,x,t)&=&\rho\int_{-\infty}^0\!\! ds \ s V^\delta\left(s,\eta-\frac{s}{|x|},x,t\right),\\ F^\delta_3(\eta,x,t)&=&-2\rho \int_{-\infty}^0\!\! ds \ s\int_0^1d\beta\ V^\delta \left(s,\eta-\frac{\beta s}{|x|},x,t\right). \end{veqnarray} Here $V^\delta(s,\eta,x,t)=(V^\delta_+(s,\eta,x,t)+V^\delta_-(s,\eta,x,t))/2$, $V^\delta_\pm(s,\eta,x,t)$ are the PDFs of \begin{veqnarray} \nonumber (s(y_0,t),\eta_\pm(y_0,x,t)), \end{veqnarray} with $\eta_\pm(y_0,x,t)=\pm(u(y_0\pm|x|,t)-u_\pm(y_0,t))/|x|$, conditional on the property that $y_0$ is a shock position. Define \begin{veqnarray} \nonumber Q(\xi,t)=\lim_{x\to0}Q^\delta(\xi,x,t). \end{veqnarray} It is easy to see that \begin{veqnarray} \nonumber &&\int d\eta\ F^\delta_1(\eta,x,t) = \rho\< s\>,\\ &&\int d\eta \ F^\delta_2(\eta,x,t) = \rho\int d\eta \int_{-\infty}^0\!\! ds \ s V^\delta(s,\eta,x,t)=\rho\< s\>,\\ &&\int d\eta\ F^\delta_3(\eta,x,t) = -2\rho\int d\eta \int_{-\infty}^0\!\! ds \ s V^\delta(s,\eta,x,t)=-2\rho\< s\>, \end{veqnarray} consistent with the fact that \begin{veqnarray} \nonumber \int d\eta \ F^\delta(\eta,x,t) =0, \end{veqnarray} and, hence, \begin{veqnarray} \nonumber \frac{d}{dt}\int d\eta\ Q^\delta(\eta,x,t)=0, \qquad \int d\eta\ \eta Q^\delta(\eta,x,t)=0. \end{veqnarray} In the limit as $x\to0$, $\eta_+$ and $\eta_-$ converge respectively to the gradient of the velocity at the left and right sides of the shock, and, pointwise in $\eta$, we have \begin{veqnarray} \nonumber \lim_{x\to0}F^\delta_2 (\eta,x,t)=\lim_{x\to0}F^\delta_3 (\eta,x,t)=0, \end{veqnarray} \begin{veqnarray} \nonumber \lim_{x\to0}F^\delta_1 (\eta,x,t)=F (\eta,t), \end{veqnarray} with $F$ given by (\ref{eq:4.2}). Therefore, a standard argument with test functions applied to (\ref{eq:4.3}) shows that in the limit as $x\to0$, $Q^\delta$ converges weakly to $Q$, solution of (\ref{eq:4.1}). \end{proof} \demo{Remark} $F^\delta_2$ and $F^\delta_3$ are examples of what we will call ``ghost terms'', i.e. terms that have finite total moments, but in the limit converge pointwise to zero. Due to the ghost terms, is not clear at the moment that $Q$ satisfies $\int d\xi \,Q(\xi,t)=1$. This will be established as a consequence of Lemma~\ref{lem:1}. \enddemo \subsection{Alternative limiting processes} \label{sec:4.2} Using BV-calculus, one works at $\nu=0$ and access the statistics of the velocity gradient by taking $x\to0$ in $(u(x+y,t)-u(y,t))/x$. This procedure gives an equation for \begin{veqnarray} \nonumber Q(\xi,t)=\lim_{x\to0} Q^\delta(\xi,x,t), \end{veqnarray} where $Q^\delta(\xi,x,t)$ is the PDF of $(u(x+y,t)-u(y,t))/x$. In this section, we revert the order of the limits: we take $x\to0$ first, working at finite $\nu$, and then let $\nu\to0$. As in Sec.~\ref{sec:3.2}, this is done using boundary layer analysis and matched asymptotics. In this way, we will obtain an equation for \begin{veqnarray} \nonumber \underline{Q}(\xi,t)=\lim_{\nu\to0} Q^\nu(\xi,t), \end{veqnarray} which will turn out to be identical to the one for $Q$. To the extend that boundary layer analysis can be justified, this strongly suggests that the limits $x\to0$, $\nu\to0$ commute. For statistically homogeneous situations, recall that $Q^\nu(\xi,t)$ satisfies (see (\ref{eq:A.6}) in the Appendix) \begin{veqnarray} \nonumber Q^\nu_t=\xi Q^\nu +\bigl(\xi^2 Q^\nu\bigr)_\xi +B_1 Q^\nu_{\xi\xi} -\nu\bigl( \< \xi_{xx}|\xi\> Q^\nu\bigr)_\xi. \end{veqnarray} The average of the dissipative term can be expressed as \begin{veqnarray} \nonumber \nu \< \xi_{xx}|\xi\> Q^\nu = \nu\lim_{L\to\infty}\frac{1}{2L} \int_{-L}^{L} dx\ \xi_{xx}(x,t) \delta(\xi-\xi(x,t)). \end{veqnarray} As in Sec.~\ref{sec:3.2}, we will evaluate this integral using the approximation for $\xi_{xx}$ provided by the boundary layer analysis. However, this analysis has to be developed further in order to evaluate the dissipative term for the velocity gradient. Recall that $\xi^\inn = u^\inn_x = {v_0}_z/\nu+{v_1}_z+O(\nu)$. Inside the shock, the $O(1)$ contribution of ${v_1}_z$ is clearly negligible compared to the $O(\nu^{-1})$ contribution of ${v_0}_z/\nu$. However, the contribution of ${v_1}_z$ {\em is} important at the border of the shock because ${v_0}_z$ decays exponentially fast there. To evaluate ${v_1}_z$ at the shock boundaries $z\to\pm\infty$, consider the equation for $v_1$ that we get from (\ref{eq:3.13}): \begin{veqnarray} \nonumber {v_0}_t+(v_0-\bar u){v_1}_z+{v_0}_z(v_1+\gamma) = {v_1}_{zz}+f. \end{veqnarray} The general solution of this equation can be expressed as \begin{veqnarray} \nonumber v_1 &=& C_1 {\rm e}^{\int_0^z\! dz'\, (v_0(z')-\bar u)}\\ &+& \int_0^z\!\!dz' \biggl(C_2+(v_0(z')-\bar u)\gamma+ \int_0^{z'}dz'' {v_0}_t(z'')-f z'\biggr) {\rm e}^{\int_{z'}^z\! dz''\, (v_0(z'')-\bar u)}, \end{veqnarray} where $C_1$, $C_2$ are constants. They, as well as $\gamma$, have to be determined by matching with $u^{\hbox{\tiny out}}} \def\inn{{\hbox{\tiny in}}=u_0+u_1+O(\nu^2)$ \cite{goxi92}. We will not dwell on this problem since $C_1$, $C_2$, $\gamma$ do not enter the expression for ${v_1}_z$ as $z\to\pm\infty$. Indeed, using the expression for $v_0$, direct computation shows that $v_1$ reduces asymptotically to \begin{veqnarray} \nonumber v_1 &=& \frac{1}{s} \biggl(2\frac{d\bar u}{dt}-\frac{ds}{dt} -2f\biggr) z\\ &+&\frac{2}{s^2}\biggl(C_2s+\frac{ds}{dt} -2\frac{d\bar u}{dt}-\frac{s^2\gamma}{2}+2f\biggr) +O\left({\rm e}^{-sz/2}\right)\qquad \hbox{as } z\to-\infty, \end{veqnarray} and \begin{veqnarray} \nonumber v_1&=& -\frac{1}{s} \biggl(2\frac{d\bar u}{dt}+\frac{ds}{dt} -2f\biggr) z\\ &-&\frac{2}{s^2}\biggl(C_2s+\frac{ds}{dt}+2\frac{d\bar u}{dt} +\frac{s^2\gamma}{2}-2f\biggr)+O\left({\rm e}^{sz/2}\right)\qquad \hbox{as } z\to+\infty. \end{veqnarray} Thus \begin{veqnarray} \nonumber \lim_{z\to\pm\infty} {v_1}_z = \mp \frac{2}{s}\frac{d\bar u}{dt} -\frac{1}{s} \frac{ds}{dt} \pm\frac{2f}{s}\equiv\xi_\pm, \end{veqnarray} where the last equality is simply a definition of $\xi_\pm$. Note that these can be reorganized to give \begin{veqnarray} \nonumber \frac{ds}{dt} = -\frac{s}{2} \bigl( \xi_+ +\xi_-\bigr), \qquad \frac{d\bar u}{dt} = -\frac{s}{4} \bigl( \xi_+ -\xi_-\bigr)+f. \end{veqnarray} In the limit as $\nu\to0$ these are the equations of motion along the shock. Using these results, to $O(\nu)$, (\ref{eq:3.12}) can be estimated as \begin{veqnarray} \nonumber \nu \< \xi_{xx} |\xi\> Q^\nu &=& \nu \rho \int ds d\bar ud\xi_+ d\xi_- \ S (\bar u,s,\xi_+,\xi_-,t) \\ &&\qquad \times \int_{\Omega} dx\ \xi^\inn_{xx}(x,t) \delta(\xi-\xi^\inn(x,t)), \end{veqnarray} where $\Omega$ is a layer centered at $y$ with width $\gg O(\nu)$, $S(\bar u,s,\xi_+,\xi_-,t)$ is the PDF of $(\bar u(y_0,t),s(y_0,t),\xi_+(y_0,t),\xi_-(y_0,t))$, conditional on $y_0$ being a shock location ($\xi_+$ and $\xi_-$ have to be included for reasons to be made clear later). We now go to the stretched coordinate $z=(x-y)/\nu$, and use the result of Sec.~\ref{sec:3.2}, namely, \begin{veqnarray} \nonumber \label{eq:xiin} \xi^\inn = \nu^{-1} {v_0}_z +{v_1}_z +O(\nu). \end{veqnarray} To $O(\nu)$, both terms must be included. However the contribution of ${v_1}_z$ is important only at the border of the shock where ${v_0}_z$ falls exponentially fast, and can be neglected inside the shock. Thus, $\xi^\inn \approx \bar \xi= \nu^{-1} {v_0}_z +\xi_\pm$, and we have \begin{veqnarray} \nonumber \nu \< \xi_{xx} |\xi\> Q^\nu &=&\rho \int ds d\bar ud\xi_+ d\xi_- \ S (\bar u,s,\xi_+,\xi_-,t)\\ &&\qquad \times\int_{-\infty}^{+\infty}\!\! dz \ \bar\xi_{zz}(z,t) \delta(\xi-\bar\xi(z,t)). \end{veqnarray} To perform the $z$ integral, we change the integration variable to $\xi'=\bar \xi-\xi_-$ for $z<0$ and to $\xi'=\bar \xi-\xi_+$ for $z>0$. For $z<0$ \begin{veqnarray} \nonumber &&((v_0-\bar u)(\bar \xi-\xi_-))_z=\bar \xi_{zz},\\ &&(v_0-\bar u)=\left({\t\frac{1}{4}}s^2+2\nu(\bar \xi-\xi_-)\right)^{1/2}= -\frac{s}{2} +O(\nu), \end{veqnarray} and we have \begin{veqnarray} \nonumber dz {\bar \xi}_{zz}=d\xi'\ \frac{\bar \xi_{zz}}{(\bar \xi-\xi_\pm)_z}=- d\xi'\ \frac{(s/2(\bar \xi-\xi_\pm))_z}{(\bar \xi-\xi_\pm) _z} = -d\xi' \frac{s}{2}. \end{veqnarray} Similarly, for $z>0$ \begin{veqnarray} \nonumber &&((v_0-\bar u)(\bar \xi-\xi_+))_z=\bar \xi_{zz},\\ &&(v_0-\bar u)=-\left({\t\frac{1}{4}}s^2+2\nu(\bar \xi-\xi_+)\right)^{1/2}= \frac{s}{2} +O(\nu), \end{veqnarray} and we have \begin{veqnarray} \nonumber dz {\bar \xi}_{zz}=d\xi'\ \frac{\bar \xi_{zz}}{(\bar \xi-\xi_\pm)_z}= d\xi'\ \frac{(s/2(\bar \xi-\xi_\pm))_z}{(\bar \xi-\xi_\pm) _z} = d\xi' \frac{s}{2}. \end{veqnarray} This leads to \begin{veqnarray} \nonumber &&\nu \< \xi_{xx}|\xi\> Q^\nu\\ &=&\rho \int ds d\bar ud\xi_+ d\xi_- \ S (\bar u,s,\xi_+,\xi_-,t)\int^{0}_{-s^2/(8\nu)} \!\! d \xi'\ \frac{s}{2}\, \delta(\xi-\xi'-\xi_-)\\ &\ns{+}&\rho \int ds d\bar ud\xi_+ d\xi_- \ S (\bar u,s,\xi_+,\xi_-,t)\int_{-s^2/(8\nu)}^{0} \!\! d\xi'\ \frac{s}{2}\, \delta(\xi-\xi'-\xi_+). \end{veqnarray} Letting $\nu\to0$ and integrating gives \begin{veqnarray} \nonumber \lim_{\nu\to0}\nu \< \xi_{xx}|\xi\> Q^\nu= \rho\int_{-\infty}^0\!\! ds\ s \int_\xi^{\infty}\!\! d\xi'\ V(s,\xi',t), \end{veqnarray} where we used the consistency constraint \begin{veqnarray} \nonumber \int d\bar u d\xi_\mp\, S (\bar u,s,\xi_+,\xi_-,t) =V_\pm(s,\xi_\pm,t). \end{veqnarray} Thus \begin{veqnarray} \nonumber -\lim_{\nu\to0}\nu \bigl(\< \xi_{xx}|\xi\> Q^\nu\bigr)_\xi= \rho\int_{-\infty}^0\!\! ds\ s V(s,\xi,t)=F(\xi,t). \end{veqnarray} \subsection{Consequences of the master equation} \label{sec:4.3} First we observe \begin{lemma} \label{lem:1} \begin{veqnarray} \label{eq:4.3.1} \< \xi\> + \rho \< s\>=0, \quad \hbox{or }\quad \int d\xi \ \xi Q(\xi,t) =-\rho \< s\>. \end{veqnarray} \end{lemma} \begin{proof} Denote by $u^\nu$ the solution of (\ref{eq:1.1}) for finite $\nu$. Then $\< u^\nu_x\>=0$. Let $\varphi(x)$ be a compactly supported smooth function. We have \begin{veqnarray} \nonumber 0=\int dx \ \varphi \< u^\nu_x\> =\Bigl\< \int dx \ \varphi u^\nu_x\Bigr\> =-\Bigl\< \int dx \ \varphi_x u^\nu\Bigr\>. \end{veqnarray} In the limit as $\nu\to0$, $u^\nu\to u$, the solution of (\ref{eq:3.3}). Thus \begin{veqnarray} \nonumber 0= -\Bigl\< \int dx \ \varphi_x u\Bigr\>. \end{veqnarray} Denote by $y_j$ the location of the shocks. We can write \begin{veqnarray} \nonumber \int dx \ \varphi_x u=-\sum_j \int_{y_j}^{y_{j+1}}\!\! dx \ \varphi u_x-\sum_j \varphi(y_j) s(y_j,t). \end{veqnarray} Averaging this result, we get \begin{veqnarray} \nonumber \int dx \ \varphi \bigl(\< \xi\> +\rho\< s\>\bigr) = 0, \end{veqnarray} for all compactly supported smooth functions $\varphi$. Hence (\ref{eq:4.3.1}). \end{proof} Notice that for finite $\nu$ or $x$, $\int d\xi\, \xi Q^\nu(\xi,t)= \int d\xi\, \xi Q^\delta(\xi,x,t)=0$. In the language of Kraichnan \cite{kra99}, (\ref{eq:4.3.1}) represents a flow of probability of $\xi$ from the smooth part of the velocity field to the shocks. It reflects the fact that no matter how small $\nu$ is, the dissipation range has a finite effect on inertial range statistics. As a consequence of (\ref{eq:4.1}) and (\ref{eq:4.3.1}), we have \begin{veqnarray} \nonumber \frac{d}{dt} \int d\xi \ Q(\xi,t)= 0, \end{veqnarray} i.e. the normalization of $Q$ is preserved. For the initial data we are interested in, this implies \begin{veqnarray} \nonumber \int d\xi \ Q(\xi,t)= 1. \end{veqnarray} \begin{lemma} \label{lem:2} \begin{veqnarray} \label{eq:4.5} \frac{d}{dt}(\rho\< s\>)=- \frac{\rho}{2}(\< s\xi_+\>+\< s\xi_-\>). \end{veqnarray} In particular $\int d\xi \, \xi F(\xi,t) = \rho(\< s\xi_+\>+\< s\xi_-\>)/2$ exists and is finite. \end{lemma} \begin{proof} Since $\rho\< s\>$ and its derivative are obviously finite, the existence of the integral $\int d\xi \, \xi F(\xi,t) = \rho(\< s\xi_+\>+\< s\xi_-\>)/2$ follows from the argument below and a standard cut-off argument on large values of $\xi_\pm$. Using \begin{veqnarray} \nonumber {u_\pm}_t+u_\pm \xi_\pm=f, \end{veqnarray} where $\xi_\pm(x,t)=u_x(x\pm,t)$, and $dy_j/dt=\bar u(y_j,t)$, we have \begin{veqnarray} \nonumber \frac{d}{dt} u_+(y_j,t)&=& \frac{dy_j}{dt} \xi_+(y_j,t) +{u_+}_t(y_j,t)\\ &=& \bar u(y_j,t)\xi_+(y_j,t) - u_+(y_j,t)\xi_+(y_j,t)+f(y_j,t)\\ &=& -\frac{1}{2}s(y_j,t)\xi_+(y_j,t) + f(y_j,t). \end{veqnarray} Similarly \begin{veqnarray} \nonumber \frac{d}{dt} u_-(y_j,t)=\frac{1}{2}s(y_j,t)\xi_-(y_j,t) + f(y_j,t). \end{veqnarray} These equations can be reorganized into \begin{veqnarray} \nonumber \frac{d}{dt} \bar u(y_j,t)= -\frac{1}{4}s(y_j,t)(\xi_+(y_j,t)-\xi_-(y_j,t))+f(y_j,t). \end{veqnarray} and \begin{veqnarray} \nonumber \frac{d}{dt} s(y_j,t)=- \frac{1}{2}s(y_j,t)(\xi_+(y_j,t)+\xi_-(y_j,t)). \end{veqnarray} Consider now \begin{veqnarray} \label{eq:4.6} \rho\< s\> = \biggl\< \sum_j s(x,t) \delta(x-y_j)\biggr\>= \biggl\< \sum_j s(y_j,t) \delta(x-y_j)\biggr\>. \end{veqnarray} Using the equation for $s(y_j,t)$ and $dy_j/dt=\bar u(y_j,t)$, we have \begin{veqnarray} \nonumber \frac{d}{dt} (\rho\< s\>)&=&- \frac{\rho}{2}(\< s\xi_+\>+\< s\xi_-\>)\\ &&-\biggl\< \sum_j s(y_j,t)\bar u(y_j,t) \delta(x-y_j)\biggr\>_x\\ && +\hbox{contribution from shock creation and collision}. \end{veqnarray} The second term at the right hand-side is zero by homogeneity. Thus to obtain (\ref{eq:4.5}) it remains to prove that the contribution of shock creation and collision vanish. To understand how this term arises and why it vanishes, assume a shock is created at position $y_1$ at time $t_1$. Then the sum under the average in (\ref{eq:4.6}) involves a term like \begin{veqnarray} \nonumber T_1=s(y_1,t)\delta(x-y_1) H(t-t_1). \end{veqnarray} where $H(\cdot)$ is the Heaviside function. Time differentiation gives \begin{veqnarray} \nonumber \frac{dT_1}{dt}&=&\frac{d}{dt}s(y_1,t) \delta(x-y_1) H(t-t_1)- s(y_1,t)\bar u(y_1,t)\delta_x(x-y_1) \\ &+& s(y_1,t)\delta(x-y_1) \delta(t-t_1). \end{veqnarray} The last term accounts for shock creation. Since the shock amplitude is zero at creation, \begin{veqnarray} \nonumber s(y_1,t)\delta(x-y_1) \delta(t-t_1) =s(y_1,t_1)\delta(x-y_1) \delta(t-t_1)=0. \end{veqnarray} This means that shock creation makes no contribution to the time derivative of (\ref{eq:4.6}). Consider now the collision events. Assume at time $t_1$ the shocks located at $y_2$ and $y_3$ merge into one shock located at $y_1$. Obviously $y_1(t_1)=y_2(t_1)=y_3(t_1)$. Such an event contributes to the sum under the average in (\ref{eq:4.6}) by a term like \begin{veqnarray} \nonumber T_2&=&s(y_1,t)\delta(x-y_1) H(t-t_1)\\ &+&\bigl(s(y_2,t)\delta(x-y_2)+s(y_3,t)\delta(x-y_3)\bigr)H(t_1-t). \end{veqnarray} Time differentiation gives \begin{veqnarray} \nonumber \frac{dT_2}{dt}&=& \frac{d}{dt}s(y_1,t)\delta(x-y_1) H(t-t_1)\\ &+&\frac{d}{dt}s(y_2,t)\delta(x-y_2) H(t_1-t) +\frac{d}{dt}s(y_3,t)\delta(x-y_3)H(t_1-t)\\ &-& s(y_1,t)\bar u(y_1,t) \delta_x(x-y_1) H(t-t_1)\\ &-&\bigl(s(y_2,t)\bar u(y_2,t)\delta_x(x-y_2) +s(y_3,t)\bar u(y_3,t)\delta_x(x-y_3)\bigr)H(t_1-t)\\ &+& \bigl(s(y_1,t)\delta(x-y_1) -s(y_2,t)\delta(x-y_2)-s(y_3,t)\delta(x-y_3)\bigr) \delta(t-t_1). \end{veqnarray} The term involving $\delta(t-t_1)$ arises from shock collision. Since $y_1(t_1)=y_2(t_1)=y_3(t_1)$ and shock amplitudes add up at collision, \begin{veqnarray} \nonumber &&\lim_{t\to0+}s(y_1(t_1+t),t_1+t)\delta(x-y_1(t_1+t))\\ &=&\lim_{t\to0+} \bigl(s(y_2(t_1-t),t_1-t)\delta(x-y_2(t_1-t))\\ &&\quad\ \, +s(y_3(t_1-t),t_1-t) \delta(x-y_3(t_1-t))\bigr), \end{veqnarray} the term in the equation for $T_2$ involving $\delta(t-t_1)$ vanishes. This means shock collision makes no contribution to the time derivative of (\ref{eq:4.6}). Hence (\ref{eq:4.5}). \end{proof} \begin{corollary} \label{lem:4} We have \begin{veqnarray} \label{eq:4.8} \lim_{|\xi|\to\infty} |\xi|^3 Q(\xi,t)=0, \end{veqnarray} i.e. $Q$ decays faster than $|\xi|^{-3}$ as $\xi\to-\infty$. \end{corollary} \begin{proof} Taking the first moment of (\ref{eq:4.1}) leads to \begin{veqnarray} \nonumber \frac{d}{dt}\< \xi\> =\left[\xi^3Q\right]_{-\infty}^{+\infty} + \int d\xi \ \xi F=\left[\xi^3Q\right]_{-\infty}^{+\infty} + \frac{\rho}{2} \left(\< s\xi_+\>+\< s\xi_-\>\right) \end{veqnarray} Using (\ref{eq:4.3.1}) this equation can be written as \begin{veqnarray} \nonumber \frac{d}{dt}\bigl(\rho\< s\> \bigr) =-\left[\xi^3Q\right]_{-\infty}^{+\infty} - \frac{\rho}{2} \left(\< s\xi_+\>+\< s\xi_-\>\right). \end{veqnarray} Using (\ref{eq:4.5}) we obtain that the boundary term must be zero. Since $\xi^3Q$ has different sign for large negative and positive values of $\xi$, one must have separately $\lim_{\xi\to-\infty}|\xi|^3Q=0$ and $\lim_{\xi\to+\infty}|\xi|^3Q=0$. Hence (\ref{eq:4.8}). \end{proof} \demo{Remark} Lemma \ref{lem:2} uses the fact that shocks are created at zero amplitude and shock strengths add up at collision, which holds in the case of large-scale smooth forcing \cite{ekhma99}. It is possible to construct pathological situations, such as the unforced case with piecewise linear initial data \cite{pol99}, in which case shocks are created at finite amplitude. In this case Lemma \ref{lem:2} is changed to \begin{lemma} \label{lem:3} \begin{veqnarray} \label{eq:4.7} \frac{d}{dt}(\rho\< s\>)=- \frac{\rho}{2}(\< s\xi_+\>+\< s\xi_-\>)-D_c, \end{veqnarray} where \begin{veqnarray} \nonumber D_c=-\sigma_1 \< s_1\>-\sigma_2 \< s_2\>\ge0. \end{veqnarray} Here $\sigma_1$ and $\sigma_2$ are respectively the space-time number density of shock creation and collision points, $\< s_1\>\le0$ is the average shock amplitude at creation, and $\< s_2\>\le0$ is the average gain of amplitude at shock collision. In this case \begin{veqnarray} \label{eq:4.8.1} Q(\xi,t)\sim D_c\, |\xi|^{-3}\qquad \hbox{as } \xi\to-\infty. \end{veqnarray} \end{lemma} Lemma \ref{lem:3} follows from a direct adaptation of the proof of Lemma \ref{lem:2}. \enddemo \subsection{Realizability and asymptotics for the statistical steady state} \label{sec:4.4} We now turn to the study of (\ref{eq:4.1}) at steady state \begin{veqnarray} \label{eq:4.4.1} 0=\xi Q +\bigl(\xi^2 Q\bigr)_\xi+ B_1 Q_{\xi\xi}+F(\xi), \end{veqnarray} where $F(\xi)=\lim_{t\to\infty} F(\xi,t)$. We shall prove that \begin{theorem} \label{th:4.6} The realizability constraint $Q\in L^1(\hbox {\Bbb R}} \def\N{\hbox {\Bbb N})$, $Q\ge0$ implies \begin{veqnarray} \label{eq:4.4.2} \lim_{\xi\to+\infty} \xi^{-2} {\rm e}^{\Lambda} F(\xi)=0, \end{veqnarray} where $\Lambda=-\xi^3/3B_1$. Assuming (\ref{eq:4.4.2}), the only positive solution of (\ref{eq:4.4.1}) can be expressed as \begin{veqnarray} \label{eq:4.4.3} Q_\infty(\xi)= \frac{1}{B_1}\int_{-\infty}^\xi \!\! d\xi' \ \xi' F(\xi')-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi'), \end{veqnarray} where \begin{veqnarray} \nonumber G(\xi)=F(\xi)+\frac{\xi}{B_1} \int_{-\infty}^{\xi}\!\!d\xi'\ \xi' F(\xi'). \end{veqnarray} Furthermore \begin{veqnarray} \label{eq:4.4.4} Q_\infty(\xi)\sim \left\{ \begin{array}{ll} {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle |\xi|^{-3} \int_{-\infty}^\xi\!\! d\xi' \xi' F(\xi') \qquad}&\hbox{as } \xi\to-\infty,\\ {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle C_+ \xi{\rm e}^{-\Lambda}} &\hbox{as }\xi\to+\infty, \end{array}\right. \end{veqnarray} where \begin{veqnarray} \nonumber C_+=-\frac{1}{B_1}\int d\xi\ {\rm e}^{\Lambda} G(\xi). \end{veqnarray} \end{theorem} Under the conditions of Theorem~\ref{th:4.6} we also have \begin{lemma} \label{th:4.7} $Q_\infty$ can be expressed as \begin{veqnarray} \label{eq:4.4.5} Q_\infty(\xi)= -\frac{\xi{\rm e}^{-\Lambda}}{B_1}\int_{-\infty}^\xi \!\! d\xi' \ \xi'^{-2}{\rm e}^{\Lambda'} \int_{-\infty}^{\xi'}\!\!d\xi'' \ \xi'' F(\xi''), \end{veqnarray} for $\xi<0$, and \begin{veqnarray} \label{eq:4.4.6} Q_\infty(\xi)=C_+\xi{\rm e}^{-\Lambda} -\frac{\xi{\rm e}^{-\Lambda}}{B_1}\int_\xi^{+\infty} \!\! d\xi' \ \xi'^{-2}{\rm e}^{\Lambda'} \int_{\xi'}^{+\infty}\!\! d\xi'' \ \xi'' F(\xi''), \end{veqnarray} for $\xi>0$. \end{lemma} \demo{Remark 1} $\xi F(\xi)$ is integrable as a consequence of Lemma~\ref{lem:2}. In particular (\ref{eq:4.5}) at steady state gives \begin{veqnarray} \nonumber \int d\xi\ \xi F=\frac{\rho}{2}(\< s\xi_+\>+\< s\xi_-\>)=0. \end{veqnarray} \enddemo \demo{Remark 2} $C_+$ is finite if (\ref{eq:4.4.2}) holds. To see this, note that because of the factor ${\rm e}^{\Lambda}$, problem may arise at $\xi=+\infty$ only. (\ref{eq:4.4.2}) implies that we can write \begin{veqnarray} \nonumber F(\xi)= {\rm e}^{-\Lambda} f(\xi), \end{veqnarray} with $f(\xi)=o(\xi^2)$ as $\xi\to+\infty$. Using $\int d\xi\, \xi F=0$, we write $G(\xi)$ as \begin{veqnarray} \nonumber G(\xi)&=&F(\xi)-\frac{\xi}{B_1} \int_{\xi}^{+\infty}\!\!d\xi'\ \xi' F(\xi')\\ &=&{\rm e}^{-\Lambda}f(\xi)-\frac{\xi}{B_1} \int_{\xi}^{+\infty}\!\!d\xi'\ \xi' {\rm e}^{-\Lambda'} f(\xi'). \end{veqnarray} Using ${\rm e}^{-\Lambda'}= -B_1 \xi'^{-2} ({\rm e}^{-\Lambda'})_{\xi'}$, after integration by parts we have \begin{veqnarray} \nonumber G(\xi)=-\xi\int_{\xi}^{+\infty}\!\!d\xi'\ (\xi'^{-1}f(\xi'))_{\xi'} {\rm e}^{-\Lambda'}. \end{veqnarray} Since $f(\xi)=o(\xi^2)$ as $\xi\to+\infty$, $(\xi^{-1}f(\xi))_{\xi}=o(1)$ and \begin{veqnarray} \nonumber G(\xi)=\int_{\xi}^{+\infty}\!\!d\xi'\ o(1) {\rm e}^{-\Lambda'} =o\left(\int_{\xi}^{+\infty}\!\!d\xi'\ {\rm e}^{-\Lambda'}\right) =o\left(\xi^{-2}{\rm e}^{-\Lambda}\right), \end{veqnarray} where at the last step we used ${\rm e}^{-\Lambda}=O((\xi^{-2}{\rm e}^{-\Lambda})_\xi)$. Thus \begin{veqnarray} \nonumber {\rm e}^{\Lambda}G(\xi)=o(\xi^{-2}) \qquad \hbox{as } \ \xi\to+\infty, \end{veqnarray} which implies that $C_+$ is finite. Similarly, we can show that the last integral in (\ref{eq:4.4.6}) is finite {\em iff} (\ref{eq:4.4.2}) holds. Indeed, using $F(\xi)=o(\xi^2{\rm e}^{-\Lambda})$ as $\xi\to+\infty$ we have \begin{veqnarray} \nonumber \int_{\xi'}^{+\infty}\!\! d\xi'' \ \xi'' F(\xi'') &=& \int_{\xi'}^{+\infty}\!\! d\xi''\ o(\xi''^3{\rm e}^{-\Lambda''})\\ &=& o\Bigl(\int_{\xi'}^{+\infty}\!\! d\xi''\ \xi''^3{\rm e}^{-\Lambda''}\Bigr)\\ &=& o(\xi'{\rm e}^{-\Lambda'}), \end{veqnarray} where we used $\xi^3{\rm e}^{-\Lambda}=O((\xi{\rm e}^{-\Lambda})_\xi)$. Thus \begin{veqnarray} \label{eq:4.4.6.b} \xi'^{-2}{\rm e}^{\Lambda'}\int_{\xi'}^{+\infty}\!\! d\xi'' \ \xi'' F(\xi'')= o(\xi'^{-1}) \qquad \hbox{as } \ \xi\to+\infty, \end{veqnarray} which is the necessary and sufficient condition for the last integral in (\ref{eq:4.4.6}) to be finite. \enddemo \begin{proof}[Proof of Lemma \ref{th:4.7}] To show (\ref{eq:4.4.5}), we start from the explicit expression for the integral involving $G$ in (\ref{eq:4.4.3}) \begin{veqnarray} \nonumber &&-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')\\ &=& -\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} F(\xi') -\frac{\xi{\rm e}^{-\Lambda}}{B^2_1} \int_{-\infty}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} \xi'\int_{-\infty}^{\xi'}\!\! d\xi'' \ \xi'' F(\xi''), \end{veqnarray} and integrate by parts the second integral using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{-\Lambda'})_{\xi'}$. The result is \begin{veqnarray} \nonumber &&-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')\\ &=& -\frac{1}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ \xi' F(\xi') -\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ \xi'^{-2} {\rm e}^{\Lambda'} \int_{-\infty}^{\xi'}\!\! d\xi'' \ \xi'' F(\xi''). \end{veqnarray} Inserting this expression in (\ref{eq:4.4.3}) gives (\ref{eq:4.4.5}). To show (\ref{eq:4.4.6}), we use $\int d\xi \xi F=0$ and write (\ref{eq:4.4.3}) as \begin{veqnarray} \nonumber Q_\infty(\xi)= C_+\xi{\rm e}^{-\Lambda} -\frac{1}{B_1}\int_{\xi}^{+\infty} \!\! d\xi' \ \xi' F(\xi')+\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_\xi^{+\infty} \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi'). \end{veqnarray} Using $\int d\xi \xi F=0$ we write explicitly the integral involving $G$ in this expression as \begin{veqnarray} \nonumber &&\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_\xi^{+\infty} \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')\\ &=& \frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_\xi^{+\infty} \!\! d\xi' \ {\rm e}^{\Lambda'} F(\xi') -\frac{\xi{\rm e}^{-\Lambda}}{B^2_1} \int_\xi^{+\infty} \!\! d\xi' \ {\rm e}^{\Lambda'} \xi'\int_{\xi'}^{+\infty}\!\! d\xi'' \ \xi'' F(\xi''), \end{veqnarray} and integrate by parts the second integral using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{-\Lambda'})_{\xi'}$: \begin{veqnarray} \nonumber &&\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_\xi^{+\infty} \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')\\ &=& \frac{1}{B_1} \int_\xi^{+\infty} \!\! d\xi' \ \xi' F(\xi') -\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_\xi^{+\infty} \!\! d\xi' \ \xi'^{-2} {\rm e}^{\Lambda'} \int_{\xi'}^{+\infty}\!\! d\xi'' \ \xi'' F(\xi''). \end{veqnarray} Inserting this expression in the last expression for $Q_\infty$ gives (\ref{eq:4.4.6}) \end{proof} \begin{proof}[Proof of Theorem \ref{th:4.6}] The general solution of (\ref{eq:4.4.1}) is $Q=Q_\infty +C_1 Q_1 +C_2 Q_2$, where $C_1$ and $C_2$ are constants, $Q_1$ and $Q_2$ are two linearly independent solutions of the homogeneous equation associated with (\ref{eq:4.4.1}). Two such solutions are \begin{veqnarray} \label{eq:4.4.7} Q_1(\xi)= \xi{\rm e}^{-\Lambda}, \end{veqnarray} \begin{veqnarray} \label{eq:4.4.8} Q_2(\xi)= 1- \frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ \xi'{\rm e}^{\Lambda'}, \end{veqnarray} For $\xi<0$, using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{\Lambda'})_{\xi'}$, after integration by parts $Q_2$ can be written as \begin{veqnarray} \label{eq:4.4.9} Q_2(\xi)=-\xi{\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi' \ \xi'^{-2}{\rm e}^{\Lambda'}. \end{veqnarray} We now show that realizability requires $C_1=C_2=0$. First, one readily checks that $\lim_{\xi\to-\infty} Q_\infty= \lim_{\xi\to-\infty} Q_2=0$, while $Q_1$ grows unbounded as $\xi\to-\infty$. Hence in order that $Q$ be integrable we must set $C_1=0$ and the general solution of (\ref{eq:4.4.1}) is \begin{veqnarray} \nonumber Q(\xi)=C_2 Q_2(\xi)+Q_\infty(\xi). \end{veqnarray} We evaluate this solution asymptotically as $\xi\to\pm\infty$. Consider $Q_2$ for large negative $\xi$ first. Using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{\Lambda'})_{\xi'}$, after integration by parts we write (\ref{eq:4.4.9}) as \begin{veqnarray} \nonumber Q_2(\xi) =B_1|\xi|^{-3} -4B_1\xi {\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi' \ \xi'^{-5}{\rm e}^{\Lambda'}. \end{veqnarray} Since $\xi^{-5} {\rm e} ^\Lambda=O((\xi^{-7}{\rm e} ^\Lambda)_\xi)$ as $\xi\to-\infty$, the integral in this expression is of the order \begin{veqnarray} \nonumber \xi {\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi' \ \xi'^{-5}{\rm e}^{\Lambda'} &=&\xi {\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi' O((\xi^{-7}{\rm e} ^\Lambda)_\xi)\\ &=&O\Bigl(\xi {\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi' (\xi^{-7}{\rm e} ^\Lambda)_\xi\Bigr)\\ &=&O(\xi^{-6}). \end{veqnarray} Thus \begin{veqnarray} \nonumber Q_2(\xi)=B_1|\xi|^{-3}+O(\xi^{-6})\qquad \hbox{as } \ \xi\to-\infty, \end{veqnarray} Consider now $Q_\infty$ for large negative $\xi$. Using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{\Lambda'})_{\xi'}$, after integration by parts we write (\ref{eq:4.4.5}) as \begin{veqnarray} \nonumber Q_\infty(\xi)&=&|\xi|^{-3} \int_{-\infty}^\xi d\xi'\ \xi' F(\xi')\\ &+&\xi{\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi' \Bigl(\xi'^{-4}\int_{-\infty}^{\xi'} d\xi''\ \xi''\ F(\xi'')\Bigr)_{\xi'} {\rm e}^{\Lambda'}. \end{veqnarray} As $\xi\to-\infty$ \begin{veqnarray} \nonumber \xi{\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi' \Bigl(\xi'^{-4}\int_{-\infty}^{\xi'} d\xi''\ \xi''\ F(\xi'')\Bigr)_{\xi'} {\rm e}^{\Lambda'} &=& \xi{\rm e}^{-\Lambda} \int_{-\infty}^\xi \!\! d\xi'\ o(\xi'^{-5}) {\rm e}^{\Lambda'}\\ &=& o\Bigl(\xi{\rm e}^{-\Lambda}\int_{-\infty}^\xi \!\! d\xi'\ \xi'^{-5}{\rm e}^{\Lambda'}\Bigr)\\ &=&o(\xi^{-6}), \end{veqnarray} where we used the estimate above. Thus \begin{veqnarray} \nonumber Q_\infty(\xi)=|\xi|^{-3} \int_{-\infty}^\xi d\xi'\ \xi' F(\xi') +o(\xi^{-6})\qquad \hbox{as } \ \xi\to-\infty. \end{veqnarray} Combining these expressions, we have \begin{veqnarray} \label{eq:4.4.10} Q(\xi) &=& C_2B_1|\xi|^{-3} + |\xi|^{-3} \int_{-\infty}^\xi d\xi'\ \xi' F(\xi')\\ &+&O(C_2\xi^{-6})+o(\xi^{-6})\qquad \hbox{as } \ \xi\to-\infty. \end{veqnarray} Consider now $Q_2$ for large positive $\xi$. Write (\ref{eq:4.4.8}) as \begin{veqnarray} \nonumber Q_2(\xi)= 1- \frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^{\xi_\star} \!\! d\xi' \ \xi'{\rm e}^{\Lambda'}- \frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^{\xi} \!\! d\xi' \ \xi'{\rm e}^{\Lambda'}, \end{veqnarray} where $\xi_\star >0$ is arbitrary but fixed. Using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{\Lambda'})_{\xi'}$, after integration by parts of the second integral we get \begin{veqnarray} \nonumber Q_2(\xi)&=& \xi\xi_\star^{-1}{\rm e}^{-\Lambda+\Lambda_\star} - \frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^{\xi_\star} \!\! d\xi' \ \xi'{\rm e}^{\Lambda'} - \xi{\rm e}^{-\Lambda} \int_{\xi_\star}^{\xi} \!\! d\xi' \ \xi'^{-2}{\rm e}^{\Lambda'}\\ &=&- \xi{\rm e}^{-\Lambda} \int_{\xi_\star}^{\xi} \!\! d\xi' \ \xi'^{-2}{\rm e}^{\Lambda'}+O(\xi{\rm e}^{-\Lambda}). \end{veqnarray} Integrating by parts again gives \begin{veqnarray} \nonumber Q_2(\xi)=-B_1\xi^{-3} -4B_1\xi {\rm e}^{-\Lambda} \int_{\xi_\star}^\xi \!\! d\xi' \ \xi'^{-5}{\rm e}^{\Lambda'}+O(\xi{\rm e}^{-\Lambda}). \end{veqnarray} Since $\xi^{-5} {\rm e} ^\Lambda=O((\xi^{-7}{\rm e} ^\Lambda)_\xi)$ as $\xi\to+\infty$, the remaining integral can be estimated as above, yielding \begin{veqnarray} \nonumber Q_2(\xi)=-B_1\xi^{-3} +O(\xi^{-6}). \end{veqnarray} Considering $Q_\infty$ for large positive $\xi$, we must distinguish two cases. If (\ref{eq:4.4.2}) does not hold, then in (\ref{eq:4.4.3}) we decompose \begin{veqnarray} \nonumber \frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^{\xi} \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')&=& \frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{-\infty}^{\xi_\star} \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')+\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')\\ &=&\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')+O(\xi{\rm e}^{-\Lambda}). \end{veqnarray} and write \begin{veqnarray} \nonumber Q_\infty(\xi)= \frac{1}{B_1}\int_{-\infty}^\xi \!\! d\xi' \ \xi' F(\xi')-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')+O(\xi{\rm e}^{-\Lambda}). \end{veqnarray} The second integral in this expression is explicitly \begin{veqnarray} \nonumber &&-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')\\ &=& -\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} F(\xi') -\frac{\xi{\rm e}^{-\Lambda}}{B^2_1} \int_{\xi_\star}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} \xi'\int_{-\infty}^{\xi'}\!\! d\xi'' \ \xi'' F(\xi''). \end{veqnarray} The integration by parts of the second integral using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{-\Lambda'})_{\xi'}$ gives \begin{veqnarray} \nonumber &&-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ {\rm e}^{\Lambda'} G(\xi')\\ &=& -\frac{1}{B_1} \int_{-\infty}^\xi \!\! d\xi' \ \xi' F(\xi') \\ &&-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ \xi'^{-2} {\rm e}^{\Lambda'} \int_{-\infty}^{\xi'}\!\! d\xi'' \ \xi'' F(\xi'')+O(\xi{\rm e}^{-\Lambda}). \end{veqnarray} Thus \begin{veqnarray} \nonumber Q_\infty(\xi)=-\frac{\xi{\rm e}^{-\Lambda}}{B_1} \int_{\xi_\star}^\xi \!\! d\xi' \ \xi'^{-2} {\rm e}^{\Lambda'} \int_{-\infty}^{\xi'}\!\! d\xi'' \ \xi'' F(\xi'')+O(\xi{\rm e}^{-\Lambda}). \end{veqnarray} Integrating by parts again using ${\rm e}^{\Lambda'}= B_1 \xi'^{-2} ({\rm e}^{\Lambda'})_{\xi'}$ gives \begin{veqnarray} \nonumber Q_\infty(\xi)&=&-\xi^{-3} \int_{-\infty}^\xi\!\! d\xi'\ \xi' F(\xi')\\ &+&\xi{\rm e}^{-\Lambda} \int_{\xi_\star}^\xi \!\! d\xi' \Bigl(\xi'^{-4}\int_{-\infty}^{\xi'}\!\! d\xi''\ \xi''\ F(\xi'')\Bigr)_{\xi'} {\rm e}^{\Lambda'}+O(\xi{\rm e}^{-\Lambda}). \end{veqnarray} The last integral can be estimate as for the large negative $\xi$. Using $\int d\xi\,\xi F=0$, this leads to \begin{veqnarray} \nonumber Q_\infty(\xi)=\xi^{-3} \int_\xi^{+\infty}\!\! d\xi'\ \xi' F(\xi') +o(\xi^{-6}), \end{veqnarray} and, hence, \begin{veqnarray} \label{eq:4.4.11} Q(\xi) &=& -C_2B_1\xi^{-3} +\xi^{-3} \int_\xi^{+\infty}\!\! d\xi'\ \xi' F(\xi')\\ &&+O(C_2\xi^{-6})+ o(\xi^{-6})\qquad \hbox{as } \ \xi\to+\infty. \end{veqnarray} In contrast, if (\ref{eq:4.4.2}) holds, then we can use (\ref{eq:4.4.6}). From (\ref{eq:4.4.6.b}) it follows that the integral term in this expression is $o(\xi{\rm e}^{-\Lambda})$ as $\xi\to+\infty$. Thus, \begin{veqnarray} \nonumber Q(\xi)= C_+\xi{\rm e}^{-\Lambda} +o(\xi{\rm e}^{-\Lambda}), \end{veqnarray} and \begin{veqnarray} \label{eq:4.4.12} Q_\infty(\xi)&=& -C_2B_1\xi^{-3} + C_+\xi{\rm e}^{-\Lambda}\\ &&+O(C_2\xi^{-6})+ o(\xi{\rm e}^{-\Lambda})\qquad \hbox{as } \ \xi\to+\infty. \end{veqnarray} In (\ref{eq:4.4.10}), (\ref{eq:4.4.11}), (\ref{eq:4.4.12}), if the $C_2$ term at the right-hand side is non-zero, then it will dominate the second term. However, since the $C_2$ term has opposite sign when $\xi\to\pm\infty$, it must be zero, i.e., we must set $C_2=0$. This proves that $Q=Q_\infty$. Furthermore, since the $F$ term at the right-hand side of (\ref{eq:4.4.11}) is negative (recall that $F\le0$), this solution must be rejected in order that $Q$ be non-negative. Thus (\ref{eq:4.4.2}) must hold, and we get (\ref{eq:4.4.4}). \end{proof} \demo{Remark} If the assumptions of Lemma \ref{lem:2} do not apply, i.e. if shocks are not created at zero amplitude or shock strength does not add up at collision, then from (\ref{eq:4.7}) at steady state we have \begin{veqnarray} \nonumber \int d\xi \ \xi F(\xi)= -D_c. \end{veqnarray} This implies that (\ref{eq:4.4.11}) has to be changed to \begin{veqnarray} \nonumber Q(\xi) &=& -C_2B_1\xi^{-3} +D_c\xi^{-3}+\xi^{-3} \int_\xi^{+\infty}\!\! d\xi'\ \xi' F(\xi')\\ &&+O(C_2\xi^{-6})+ o(\xi^{-6})\qquad \hbox{as } \ \xi\to+\infty. \end{veqnarray} This solution cannot be dismissed as being unrealizable since the second term at the right hand-side can balance the first one. In particular, for $C_2=D_c/B_1$ we obtain from (\ref{eq:4.4.10}) $Q(\xi)=D_c |\xi|^{-3}+o(\xi^{-3})$ as $\xi\to-\infty$, consistent with (\ref{eq:4.8.1}). \enddemo \subsection{Statistics for the environment of the shocks} \label{sec:4.5.1} We now turn to the statistics for the environment of the shocks. For simplicity we will focus on statistically homogeneous situations such that $u(x,t)\stackrel{\hbox{\tiny law}}{=}-u(-x,t)$. Define \begin{veqnarray} \nonumber s(x,y_0,t)=u(y_0+x/2,t)-u(y_0-x/2,t), \end{veqnarray} and let $W(s,\xi_+,\xi_-,x,t)$ be the PDF of \begin{veqnarray} \nonumber (s(x,y_0,t),\xi(y_0+x/2,t),\xi(y_0-x/2,t)), \end{veqnarray} conditional on $y_0$ being a shock location. Since $u(x,t)\stackrel{\hbox{\tiny law}}{=} -u(-x,t)$, it follows that \begin{veqnarray} \nonumber W(s,\xi_+,\xi_-,x,t)= W(s,\xi_-,\xi_+,x,t). \end{veqnarray} $V(s,\xi,t)$ and $W(s,\xi_+,\xi_-,x,t)$ are related by (recall that $V=V_+=V_-$ if $u(x,t)\stackrel{\hbox{\tiny law}}{=} -u(-x,t)$) \begin{veqnarray} \label{eq:4.9.l1} V(s,\xi,t)=\lim_{x\to0+} \int d\xi'\ W(s,\xi',\xi,x,t). \end{veqnarray} Thus, \begin{veqnarray} \label{eq:4.9.l2} F(\xi,t)=\rho\lim_{x\to0+}\int ds d\xi'\ W(s,\xi',\xi,x,t). \end{veqnarray} We have \begin{theorem} \label{th:3.8} $W$ satisfies \begin{veqnarray} \label{eq:4.18} (\rho W)_t&=& - \rho sW_x +2 \rho(B_0-B(x)) W_{ss}\\ &&+\frac{\rho}{2}\xi_+W +\rho\bigl(\xi_+^2 W\bigr)_{\xi_+} +\frac{\rho}{2}\xi_-W +\rho\bigl(\xi_-^2 W\bigr)_{\xi_-}\\ && + \rho B_1\bigl( W_{\xi_+ \xi_+} +W_{\xi_- \xi_-}\bigr) +2\rho B_1(x)W_{\xi_+ \xi_-}\\ &&- \rho B_2(x)\bigl( W_{s\xi_+}+W_{s \xi_-}\bigr) +\varsigma_1-\varsigma_2 +J, \end{veqnarray} with $B_1(x)=-B_{xx}(x)$, $B_2(x)=B_x(x)$. $\varsigma_1(s,\xi_+,\xi_-,x,t)$ is defined such that \begin{veqnarray} \nonumber \varsigma_1(s,\xi_+,\xi_-,x,t) ds d\xi_-d\xi_+ dz dt, \end{veqnarray} gives the average number of shock creation points in $[z,z+dz)\times[t,t+dt)$ with \begin{veqnarray} \nonumber s(x,y_1,t_1)&\in&[s,s+ds),\\ \xi(y_1+x/2,t_1)&\in&[\xi_+,\xi_++d\xi_+),\\ \xi(y_1-x/2,t_1)&\in&[\xi_-,\xi_-+d\xi_-), \end{veqnarray} conditional on $(y_1,t_1)\in ([z,z+dz)\times[t,t+dt))$ being a point of shock creation (because of the statistical homogeneity, $z$ is a dummy variable). $\varsigma_2(s,\xi_+,\xi_-,x,t)$ is defined such that \begin{veqnarray} \nonumber \varsigma_2(s,\xi_+,\xi_-,x,t) ds d\xi_-d\xi_+ dz dt, \end{veqnarray} gives the average number of shock collision points in $[z,z+dz)\times[t,t+dt)$ with \begin{veqnarray} \nonumber s(x,y_2,t_2)&\in&[s,s+ds),\\ \xi(y_2+x/2,t_2)&\in&[\xi_+,\xi_++d\xi_+),\\ \xi(y_2-x/2,t_2)&\in&[\xi_-,\xi_-+d\xi_-), \end{veqnarray} conditional on $(y_2,t_2)\in ([z,z+dz)\times[t,t+dt))$ being a point of shock collision. Finally, $J(s,\xi_+,\xi_-,x,t)$ accounts for the possibility of having another shock in between $[y_j-x/2,y_j+x/2]$ and satisfies \begin{veqnarray} \nonumber J(s,\xi_+,\xi_-,x,t)=O(x). \end{veqnarray} \end{theorem} At statistical steady state, the definitions for $\varsigma_1$, $\varsigma_2$ simplify. Indeed, in the limit as $t\to+\infty$, we have \begin{veqnarray} \nonumber \varsigma_1(s,\xi_+,\xi_-,x,t)\to \sigma_1 S_1(s,\xi_+,\xi_-,x), \end{veqnarray} where $\sigma_1$ is the space-time number density of shock creation points, $S_1(s,\xi_+,\xi_-,x)$ is the PDF of \begin{veqnarray} \nonumber (s(x,y_1,t_1),\xi(y_1+x/2,t_1),\xi(y_1-x/2,t_1)), \end{veqnarray} conditional on a shock being created at $(y_1,t_1)$, and \begin{veqnarray} \nonumber \varsigma^i_2(s,\xi_+,\xi_-,x,t)\to \sigma_2 S_2(s,\xi_+,\xi_-,x), \end{veqnarray} where $\sigma_2$ is the space-time number density of shock collision points, $S_2(s,\xi_+,\xi_-,x)$ is the PDF of \begin{veqnarray} \nonumber (s(x,y_2,t_2),\xi(y_2+x/2,t_2),\xi(y_2-x/2,t_2)), \end{veqnarray} conditional on a two shocks colliding at $(y_2,t_2)$. ~ \demo{Remark on the strategy for the proof of Theorem \ref{th:3.8}} Ideally to prove Theorem \ref{th:3.8} we should follow the strategy in Sec.~\ref{sec:4} for the derivation of the equation for $Q$. Let $X(u_1,x_1,\ldots,u_6,x_6,t)$ be the PDF of \begin{veqnarray} \nonumber (u(y_0+x_1,t),\ldots,u(y_0+x_6,t), \end{veqnarray} conditional on $y_0$ being a shock location. Knowing the equation for $X$, one can easily derive an equation for the conditional PDF of \begin{veqnarray} \nonumber (u(y_0+x/2,t),u(y_0+x/2,t),\eta(y_0+x/2,y,t),\eta(y_0-x/2,y,t) \end{veqnarray} where $\eta(x,z,t)=(u(x+z,t)-u(z,t))/z$. Letting $z\to0$, one derives an equation for $W$. Clearly, this derivation is rather tedious and, as we now show, unnecessary for our purpose. Recall that \begin{veqnarray} \nonumber \rho X(u_1,x_1,\ldots,u_6,x_6,t)&=& \int \frac{d\lambda_1\cdots d\lambda_6}{(2\pi)^6} {\rm e}^{i\lambda_1 u_1+\cdots+i\lambda_6 u_6}\\ &&\times \biggl\<\sum_j{\rm e}^{-i\lambda_1 u(z+x_1,t)-\cdots-i\lambda_6 u(z+x_1,t)} \delta(z-y_j)\biggr\>. \end{veqnarray} The average under the integral is the characteristic function associated with $X$, and an equation for this quantity can be derived using the equation for $u(z+x_p,t)$ \begin{veqnarray} \nonumber du= -{\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}} {u}_{x_p}dt +dW(z+x_p,t), \quad p=1,\ldots,6. \end{veqnarray} Instead of reproducing these straightforward calculations, we note simply that for homogeneous situations the resulting equation for $X$ will contain terms proportional to \begin{veqnarray} \nonumber \rho_2(x_p,t)=\Bigl\<\sum_{j,k} \delta(x_p+z-y_k)\delta(z-y_j)\Bigr>. \end{veqnarray} These terms account for the probability of having another shock, say $y_1$, between $y_0$ and $y_0+x_p$: they are the origin of $J$ in (\ref{eq:4.18}). Note also that technically, the $\rho_2(x_p,t)$'s arise because of the average ${\hbox{$[$}}} \def\br{{\hbox{$]$}} u\br_{\hbox{\tiny A}}} \def\B{{\hbox{\tiny B}}$ in the equation for $u(z+x_p,t)$. Now, the key point is to note that \begin{veqnarray} \nonumber \rho(x_p,t)=O(x_p). \end{veqnarray} As a direct result, \begin{veqnarray} \nonumber J(s,\xi_+,\xi_-,x,t)=O(x). \end{veqnarray} We are eventually interested in the limit as $x\to0$ of $W$. As argued in Sec.~\ref{sec:4.5}, in this limit the $O(x)$ terms in (\ref{eq:4.18}) are negligible. Thus, we will not dwell on obtaining an explicit expression for $J$. Instead, we will derive (\ref{eq:4.18}) using ($\bar W(x,t)=W_x(x,t)$) \begin{veqnarray} \label{eq:3.8.1} du&=& -u{u}_{x_\pm}dt +dW(z+x_\pm,t),\\ d\xi &=& -(u{\xi}_{x_\pm}+\xi^2)dt +d\bar W(z+x_\pm,t), \end{veqnarray} as if no shock are present between $z$ and $z+x_\pm$. The error we are making are accounted for by the term $J$. \enddemo \begin{proof}[Proof of Theorem \ref{th:3.8}] Define \begin{veqnarray} \nonumber &&\theta(\lambda_+,\lambda_-,\mu_+,\mu_-,x_+,x_-,z,t)\\ &=&{\rm e}^{-i\lambda_+u(z+x_+,t)-i\lambda_-u(z+x_-,t) -i\mu_+\xi(z+x_+,t) -i\mu_-\xi(z+x_-,t)}. \end{veqnarray} Then \begin{veqnarray} \nonumber \rho W(s,\xi_+,\xi_-,x,t) &=&\int \frac{d\lambda d\mu_+d\mu_-}{(2\pi)^3} {\rm e}^{-i\lambda s -i\mu_+\xi_+ -i\mu_-\xi_-} \\ &&\times \biggl\<\sum_j\theta(\lambda,-\lambda,\mu_+,\mu_-,x/2,-x/2,z,t) \delta(z-y_j) \biggr\>. \end{veqnarray} We now derive equations for $\<\Theta\>=\<\sum _j \theta \delta(z-y_j)\>$, $W$. We use the following rules from Ito calculus \begin{veqnarray} \nonumber d W(x,t) d W(y,t)&=&2B(x-y) dt,\\ d\bar W(x,t) d\bar W(y,t)&=&2B_1(x-y) dt\\ d W(x,t) d \bar W(y,t)&=&-2B(x-y) dt, \end{veqnarray} where $B_1(x)=-B_{xx}(x)$, $B_2(x)=B_x(x)$. Thus, from (\ref{eq:3.8.1}), we obtain \begin{veqnarray} \nonumber d\theta &=& i(\lambda_+u_+{u_+}_{x_+}+\lambda_-u_-{u_-}_{x_-} ) \theta dt\\ &-&(\lambda_+^2 B_0 +\lambda_-^2 B_0 +2 \lambda_+\lambda_- B(x_+-x_-)) \theta dt\\ &+&i(\mu_+u_+{\xi_+}_{x_+} +\mu_+\xi_+^2+\mu_-u_-{\xi_-}_{x_-}+\mu_-\xi_-^2 ) \theta dt\\ &-&(\mu_+^2 B_1 +\mu_-^2 B_1 +2 \mu_+\mu_- B_1(x_+-x_-)) \theta dt\\ &-&(\lambda_-\mu_+-\lambda_+\mu_-)B_2(x_+-x_-) \theta dt\\ &-&i\lambda_+\theta dW(z+x_+,t)-i\lambda_-\theta dW(z+x_-,t)\\ &-&i\mu_+\theta d\bar W(z+x_+,t)-i\mu_-\theta d\bar W(z+x_-,t), \end{veqnarray} where $u_\pm=u(z+x_\pm,t)$, $\xi_\pm=\xi(z+x_\pm,t)$. Similarly, using $dy_j/dt=\bar u(y_j,t)$, we get \begin{veqnarray} \nonumber d\sum_j\delta(z-y_j) &=&-\sum_j\bar u(y_j,t) \delta^1(z-y_j)dt\\ &&+\sum_k \delta(z-y_k)\delta(t-t_k)dt-\sum_l \delta(z-y_l)\delta(t-t_l)dt, \end{veqnarray} where $\delta^1(z)=d\delta(z)/dz$, the $(y_k,t_k)$'s are the points of shock creation, the $(y_l,t_l)$'s are the points of shock collisions. Using \begin{veqnarray} \nonumber \theta\delta^1(z-y_j)= (\theta\delta(z-y_j))_z-\theta_{x_+}\delta(z-y_j)- \theta_{x_-}\delta(z-y_j), \end{veqnarray} and noting that $\<\cdot\>_z=0$ by statistical homogeneity, it follows that \begin{veqnarray} \nonumber \<\Theta\>_t &=& i\lambda_+\<u_+{u_+}_{x_+}\Theta \>+ i\lambda_-\< u_-{u_-}_{x_-}\Theta \>\\ &-&(\lambda_+^2 B_0 +\lambda_-^2 B_0 +2 \lambda_+\lambda_- B(x_+-x_-)) \<\Theta\>\\ &+&i\mu_+\< (u_+{\xi_+}_{x_+} +\xi_+^2 )\Theta\> +i\mu_-\< (u_-{\xi_-}_{x_-}+\xi_-^2)\Theta\>\\ &-&(\mu_+^2 B_1 +\mu_-^2 B_1 +2 \mu_+\mu_- B_1(x_+-x_-)) \<\Theta\>\\ &-&(\lambda_-\mu_+-\lambda_+\mu_-)B_2(x_+-x_-) \<\Theta\>\\ &+&\biggl\<\sum_j\bar u(y_j,t) (\theta_{x_+}+\theta_{x_-})\delta(z-y_j)\biggr\> +{\it\Sigma}_1-{\it\Sigma}_2, \end{veqnarray} where ${\it\Sigma}_1$ and ${\it\Sigma}_2$ account respectively for shock creation and collision events. These are given by \begin{veqnarray} \nonumber &&{\it \Sigma}_1(\lambda_+,\lambda_-,\mu_+,\mu_-,x_+,x_-,t)\\ &=&\biggl\<\sum_k {\rm e}^{-i\lambda_+ u_+-i\lambda_- u_- -i\mu_+\xi_+-i\mu_-\xi_-}\delta(z-y_k)\delta(t-t_k) \biggr\>, \end{veqnarray} \begin{veqnarray} \nonumber &&{\it \Sigma}_2(\lambda_+,\lambda_-,\mu_+,\mu_-,x_+,x_-,t)\\ &=&\biggl\<\sum_l {\rm e}^{-i\lambda_+ u_+-i\lambda_- u_- -i\mu_+\xi_+-i\mu_-\xi_-}\delta(z-y_l)\delta(t-t_l) \biggr\>. \end{veqnarray} To average the convective terms we use: \begin{veqnarray} \nonumber &&i\lambda_\pm\< u_\pm{u_\pm}_{x_\pm}\Theta\> +i\mu_\pm\<(u_\pm{\xi_\pm}_{x_\pm} +\xi_\pm^2 )\Theta\>\\ &=&-\<u_\pm\Theta_{x_\pm}\> -i\mu_\pm\<\Theta\>_{\mu_\pm\mu_\pm}\\ &=&-\<u_\pm\Theta\>_{x_\pm}+ \<\xi_\pm\Theta\> -i\mu_\pm\<\Theta\>_{\mu_\pm\mu_\pm}\\ &=&-i\<\Theta\>_{x_\pm\lambda_\pm}+i\<\Theta\>_{\mu_\pm} -i\mu_\pm\<\Theta\>_{\mu_\pm\mu_\pm}. \end{veqnarray} For the term involving $\bar u(y_j,t)=(u_+(y_j,t)+u_-(y_j,t))/2$ we note that \begin{veqnarray} \nonumber u_+(y_j,t)\theta _{x_+}&=& u(y_j+x_+,t)\theta _{x_+} -x_+\int_0^1\!\! d\beta\ \xi(y_j+\beta x_+,t) \theta _{x_+}\\ &=& (u(y_j+x_+,t)\theta)_{x_+}-\xi(y_j+x_+,t)\theta\\ &&-x_+\int_0^1\!\! d\beta\ \xi(y_j+\beta x_+,t) \theta _{x_+}\\ &=& i\theta _{x_+\lambda_+}-i\theta _{\mu_+} -x_+\int_0^1\!\! d\beta\ \xi(y_j+\beta x_+,t) \theta _{x_+}. \end{veqnarray} A similar expression holds for $u_-(y_j,t)\theta _{x_-}$. Also \begin{veqnarray} \nonumber u_+(y_j,t)\theta _{x_-}&=& (u(y_j+x_+,t)\theta_j)_{x_-} -x_+\int_0^1\!\! d\beta\ \xi(y_j+\beta x_+,t) \theta _{x_-}\\ &=& i\theta _{x_-\lambda_+} -x_+\int_0^1\!\! d\beta\ \xi(y_j+\beta x_+,t) \theta _{x_-}, \end{veqnarray} and a similar expression holds for $u_-(y_j,t)\theta _{x_+}$. Thus \begin{veqnarray} \nonumber &&2\biggl\<\sum_j\bar u(y_j,t) (\theta_{x_+}+\theta_{x_-})\delta(z-y_j) \biggr\>\\ &=&i \<\Theta\>_{x_+\lambda_+}+i \<\Theta\>_{x_-\lambda_+} +i\<\Theta\>_{x_+\lambda_-}+i\<\Theta\>_{x_-\lambda_-} -i\<\Theta\>_{\mu_+} -i\<\Theta\>_{\mu_-}-R \end{veqnarray} where \begin{veqnarray} \nonumber &&R(\lambda_+,\lambda_-,\mu_+,\mu_-,x_+,x_-)\\ &=& x_+\biggl\<\sum_j \int_0^1\!\! d\beta\ \xi(y_j+\beta x_+,t) (\theta _{x_+}+\theta _{x_-})\delta(z-y_j)\biggr\>\\ &+&x_-\biggl\<\sum_j \int_0^1\!\! d\beta\ \xi(y_j+\beta x_-,t) (\theta _{x_+}+\theta _{x_-})\delta(z-y_j)\biggr\>. \end{veqnarray} Combining the above expressions leads to the following equation for~$\<\Theta\>$: \begin{veqnarray} \nonumber \<\Theta\>_t&=& -\frac{i}{2} (\<\Theta\>_{x_+\lambda_+} +\<\Theta\>_{x_-\lambda_-} -\<\Theta\>_{x_+\lambda_-}-\<\Theta\>_{x_-\lambda_+})\\ &&-(\lambda_+^2 B_0 +\lambda_-^2 B_0 +2 \lambda_+\lambda_- B(x_+-x_-)) \<\Theta\>\\ &&+\frac{i}{2}(\<\Theta\>_{\mu_+}+\<\Theta\>_{\mu_-}) -i\mu_+\<\Theta\>_{\mu_+\mu_+}-i\mu_-\<\Theta\>_{\mu_-\mu_-}\\ &&-(\mu_+^2 B_1 +\mu_-^2 B_1 +2 \mu_+\mu_- B_1(x_+-x_-)) \<\Theta\>\\ &&-(\lambda_-\mu_+-\lambda_+\mu_-)B_2(x_+-x_-) \<\Theta\> +{\it\Sigma}_1-{\it\Sigma}_2-R. \end{veqnarray} To obtain an equation for $W$, we note the following remarkable property of $R$: \begin{lemma} \begin{veqnarray} \nonumber R(\lambda_+,\lambda_-,\mu_+,\mu_-,x/2,-x/2)=0. \end{veqnarray} \end{lemma} \begin{proof} To see this, write \begin{veqnarray} \nonumber && R(\lambda_+,\lambda_-,\mu_+,\mu_-,x/2,-x/2)\\ &=&\frac{x}{2} \lim_{\bar x\to0}\frac{\partial}{\partial\bar x} \biggl\<\sum_j \int_0^1\!\! d\beta\ (\xi(y_j+\beta x/2,t)-\xi(y_j-\beta x/2,t))\theta\delta(z-y_j)\biggr\>, \end{veqnarray} where $\theta= \theta_j(\lambda_+,\lambda_-,\mu_+,\mu_-,\bar x+x/2,\bar x-x/2,z,t)$. We claim that \begin{veqnarray} \nonumber A&=&\biggl\<\sum_j \int_0^1\!\! d\beta\ (\xi(y_j+\beta x/2,t)-\xi(y_j-\beta x/2,t))\theta\delta(z-y_j)\biggr\>=0. \end{veqnarray} Indeed the symmetry $u(x,t)\stackrel{\hbox{\tiny law}}{=} -u(-x,t)$ requires that $A$ be invariant under the transformation \begin{veqnarray} \nonumber z\to-z,\quad x\to x, \quad \bar x\to -\bar x, \quad y_j\to -y_j, \quad \lambda_\pm \to -\lambda_\mp, \quad \mu_\pm \to \mu_\mp. \end{veqnarray} On the other hand one checks explicitly that $A\to-A$ under the same transformation. Hence $A=0$, $R=0$. \end{proof} We now continue with the proof of Lemma~\ref{th:3.8}. Combining the above expressions, on the subset $\lambda_+=\lambda$, $\lambda_-=-\lambda$, $x_+=x/2$, $x_-=-x/2$, $\<\Theta\>$ satisfies \begin{veqnarray} \nonumber \<\Theta\>_t&=& -i \<\Theta\>_{x\lambda}-2\lambda^2 (B_0 - B(x)) \<\Theta\>\\ &&+\frac{i}{2}(\<\Theta\>_{\mu_+}+\<\Theta\>_{\mu_-}) -i\mu_+\<\Theta\>_{\mu_+\mu_+}-i\mu_-\<\Theta\>_{\mu_-\mu_-}\\ &&-(\mu_+^2 B_1 +\mu_-^2 B_1 +2 \mu_+\mu_- B_1(x)) \<\Theta\>\\ &&+(\lambda\mu_++\lambda\mu_-)B_2(x) \<\Theta\> +{\it\Sigma}_1-{\it\Sigma}_2, \end{veqnarray} where the ${\it\Sigma}_1$, ${\it\Sigma}_2$ are evaluated at $\lambda_+=\lambda$, $\lambda_-=-\lambda$, $x_+=x/2,x_-=-x/2$. Going to the variables $(s,\xi_+,\xi_-)$ we obtain (\ref{eq:4.18}). \end{proof} \subsection{The exponent $7/2$} \label{sec:4.5} In this section we will derive the following result: ~ {\em \noindent For large negative $\xi$, $F$ behaves as \begin{veqnarray} \label{eq:4.16} F(\xi)\sim C |\xi|^{-5/2}\qquad \hbox{as } \xi\to-\infty. \end{veqnarray} } ~ A direct consequence of (\ref{eq:4.16}) is \begin{veqnarray} \label{eq:4.17} Q_\infty(\xi)\sim \left\{ \begin{array}{ll} {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle C_-|\xi|^{-7/2} \qquad}&\hbox{as } \xi\to-\infty,\\ {\displaystyle} \def\t{\textstyle} \def\scr{\scriptstyle C_+ \xi{\rm e}^{-\Lambda}} &\hbox{as }\xi\to+\infty, \end{array}\right. \end{veqnarray} The argument for (\ref{eq:4.16}) is based on the following result ~ {\em Consider \begin{veqnarray} \nonumber x^{-1} S_1(\bar s x^{1/3},\bar\xi_+ x^{-2/3},\bar\xi_-x^{-2/3},x), \end{veqnarray} the PDF of the rescaled variables \begin{veqnarray} \nonumber (s(y_1,x,t_1)x^{-1/3},\xi_+(y_1+x/2,t_1)x^{2/3},\xi_-(y_1-x/2,t_1)x^{2/3}), \end{veqnarray} conditional on a shock being created at $(y_1,t_1)$. In the limit as $x\to0$, \begin{veqnarray} \label{eq:4.19} &&x^{-1} S_1(\bar s x^{1/3},\bar\xi_+ x^{-2/3},\bar\xi_-x^{-2/3},x)\\ &&\qquad\qquad\to P(\bar s) \delta\left(\bar\xi_+-\bar s/3\right)\delta(\bar\xi_+-\bar\xi_-), \end{veqnarray} where $P(\cdot)$ is a PDF supported on $(-\infty,0]$. } ~ (\ref{eq:4.19}) shows that, in the original variables, $S_1$ is asymptotically \begin{veqnarray} \label{eq:4.19b} &&S_1(s,\xi_+,\xi_-,x)\\ &\sim& x^{-1/3} P(s x^{1/3}) \delta(\xi_+-sx^{-1}/3)\delta(\xi_+-\xi_-). \end{veqnarray} We first derive (\ref{eq:4.16}), then (\ref{eq:4.19}). \demo{Derivation of (\ref{eq:4.16})} Recall that \begin{veqnarray} \nonumber F(\xi)=\rho\lim_{x\to0+}\int ds d\xi'\ W_\infty(s,\xi',\xi,x), \end{veqnarray} where $W_\infty$ is the statistical steady state value of $W$. Thus, evaluating $F$ amounts to evaluating $W$ which we will do by analyzing (\ref{eq:4.18}). We note first that this equation describe the process of shock creation, motion, then collision. Since collision only occurs if shocks are present, it is natural to represent the effect of collision as proportional to the density of shocks in the systems, i.e. we write \begin{veqnarray} \nonumber \varsigma_2=\rho g(s,\xi_+,\xi_-,x)W, \end{veqnarray} for some function $g(s,\xi_-,\xi_+,x)$ which is assumed to be smooth in $s$, $\xi_\pm$, $x$. This amounts to assuming that the characteristics of the shocks around the collision points are not very different from the characteristics of the shocks away from the collision points. For instance, if they were identical, we would have $\varsigma_2=\sigma_2 W$ and, hence, $g=\sigma_2/\rho$. Next, since we are interested in $W$ evaluated at $x=0$, we neglect the terms \begin{veqnarray} \nonumber B_2(x)=O(x), \qquad 2(B_0-B(x))=O(x^2)\qquad J=O(x), \end{veqnarray} in (\ref{eq:4.18}). Finally, since we are interested in the limit as $\xi_\pm\to-\infty$, we neglect the forcing terms in $\xi_\pm$, proportional to $B_1$ or $B_1(x)$. Under these approximations, (\ref{eq:4.18}) reduces to \begin{veqnarray} \nonumber (\rho W)_t=- \rho sW_x +\frac{\rho}{2}(\xi_++\xi_-)W +\rho\bigl(\xi_+^2 W\bigr)_{\xi_+} +\rho\bigl(\xi_-^2 W\bigr)_{\xi_-} -\rho g W+\varsigma_1. \end{veqnarray} Assuming no shocks are present at the initial time, the equation must be solved with the initial condition $\rho W(s,\xi_+,\xi_-,x,0)=0$. The solution is \begin{veqnarray} \nonumber \rho W(s,\xi_+,\xi_-,x,t)&=&\int_0^t d\tau \ ((1-\xi_+\tau )(1-\xi_-\tau ))^{-5/2} \\ &&\times\exp\left(-\int_0^\tau \!\! d\tau '\ g\left(s,\frac{\xi_+}{1-\xi_+\tau '}% ,\frac{\xi_-}{1-\xi_-\tau '},x-\tau 's\right) \right)\\ &&\times \varsigma_1\left(s,% \frac{\xi_+}{1-\xi_+\tau },\frac{\xi_-}{1-\xi_-\tau }, x-\tau s,t-\tau \right). \end{veqnarray} The statistical steady state solution is obtained in the limit as $t\to\infty$ of this expression. Using \begin{veqnarray} \nonumber \lim_{t\to\infty} \varsigma_1(s,\xi_+,\xi_-,x,t)=\sigma_1 S_1(s,\xi_+,\xi_-,x), \end{veqnarray} we obtain \begin{veqnarray} \nonumber \rho W_\infty(s,\xi_+,\xi_-,x)&=&\sigma_1\int_0^\infty d\tau \ ((1-\xi_+\tau )(1-\xi_-\tau ))^{-5/2} \\ &&\times\exp\left(-\int_0^\tau \!\! d\tau '\ g\left(s,\frac{\xi_+}{1-\xi_+\tau '},\frac{\xi_-}{1-\xi_-\tau '},% x-\tau 's\right) \right)\\ &&\times S_1\left(s,\frac{\xi_+}{1-\xi_+\tau },\frac{\xi_-}{1-\xi_-\tau }, x-\tau s\right). \end{veqnarray} At $x=0$, using (\ref{eq:4.19b}) for $S_1$, we get \begin{veqnarray} \nonumber &&\rho W_\infty(s,\xi_+,\xi_-,0)\\ &=&\sigma_1\int_0^\infty d\tau \ ((1-\xi_+\tau )(1-\xi_-\tau ))^{-5/2} \\ &\times&\exp\left(-\int_0^\tau \!\! d\tau '\ g\left(s,\frac{\xi_+}{1-\xi_+\tau '},\frac{\xi_-}{1-\xi_-\tau '},% -\tau 's\right) \right)\\ &\times& (|s|\tau )^{-1/3} P\left(-\frac{s^{2/3}}{\tau ^{1/3}}\right) \delta\left(\frac{\xi_+}{1-\xi_+\tau }+\frac{1}{3\tau }\right) \delta\left(\frac{\xi_+}{1-\xi_+\tau }- \frac{\xi_-}{1-\xi_-\tau }\right). \end{veqnarray} Since \begin{veqnarray} \nonumber \delta\left(\frac{\xi_+}{1-\xi_+\tau }- \frac{\xi_-}{1-\xi_-\tau }\right) = (1-\xi_+\tau )^2 \delta(\xi_+-\xi_-), \end{veqnarray} we have $W_\infty(s,\xi_+,\xi_-,0)\propto \delta(\xi_+-\xi_-)$. Using the relation (\ref{eq:4.9.l1}) between $W$ and $V$, this implies that $W_\infty(s,\xi_+,\xi_-,0)=V_\infty(s,\xi_+)\delta(\xi_+-\xi_-)$, and leads to \begin{veqnarray} \nonumber \rho V_\infty(s,\xi)&=&\sigma_1\int_0^\infty d\tau \ (1-\xi \tau )^{-3} (|s|\tau )^{-1/3} P\biggl(-\frac{s^{2/3}}{\tau ^{1/3}}\biggr) \delta\left(\frac{\xi}{1-\xi \tau }+\frac{1}{3\tau }\right)\\ &&\times\exp\left(-\int_0^\tau \!\! d\tau '\ g\left(s,\frac{\xi}{1-\xi \tau '},\frac{\xi }{1-\xi \tau '},% -\tau 's\right) \right). \end{veqnarray} To perform the integration over $\tau $ we use \begin{veqnarray} \nonumber \delta\left(\frac{\xi}{1-\xi \tau }+\frac{1}{3\tau }\right)= \frac{9}{8\xi^2}\delta\left(\tau +\frac{1}{2\xi}\right). \end{veqnarray} Since we are considering $\xi\ll -1$, the exponential factor evaluated at $\tau =-1/2\xi$ is \begin{veqnarray} \nonumber \exp\left(-\int_0^{-1/2\xi}\!\! d\tau '\ g\left(s,\frac{\xi}{1-\xi \tau '},\frac{\xi }{1-\xi \tau '},% -\tau 's\right) \right)=1 +O(\xi^{-1}). \end{veqnarray} This means that shock collision events make no contribution to leading order, leaving us with \begin{veqnarray} \nonumber \rho V_\infty(s,\xi)= \bar C \sigma_1 |s| ^{-1/3} |\xi| ^{-5/3} P\left(-(2s^2|\xi|)^{1/3}\right), \end{veqnarray} where $\bar C=2^{4/3}/3$. Hence, \begin{veqnarray} \label{eq:4.19.b} F(\xi) = -\bar C \sigma_1 \int_{-\infty}^0\!\! ds \ s ^{2/3} |\xi| ^{-5/3} P\left(-(2s^2|\xi|)^{1/3}\right) = -C |\xi|^{-5/2}, \end{veqnarray} where $C = 2^{-1/2}\sigma_1 \int_{-\infty}^0 db \, |b|^{3/2} P(b)$. \enddemo ~ \demo{Derivation of (\ref{eq:4.19})} We use local analysis around the shock creation points \cite{fofr83}. Consider a shock created at $y_1$ at time $t_1$ with velocity $u_1=u(y_1,t_1)$. Assuming $x$ is an analytical function of $u$, we have locally \begin{veqnarray} \label{eq:3.31.0} x=a(u(y_1+x/2,t_1)-u_1)^3+O(u(y_1+x/2,t_1)-u_1)^4), \end{veqnarray} where $a\le0$ is a random quantity. Setting $a=(2/b)^3$ gives \begin{veqnarray} \nonumber u(y_1+x/2,t_1)&=&u_1+\frac{b}{2} x^{1/3} +O(x^{2/3}). \end{veqnarray} Hence \begin{veqnarray} \label{eq:4.9.1} s(y_1,x,t_1)&=&u(y_1+x/2,t_1)-u(y_1-x/2,t_1)\\ &=&b x^{1/3}+O(x^{2/3}), \end{veqnarray} and \begin{veqnarray} \label{eq:4.9.2} \xi(y_1+x/2,t_1)=\frac{b}{3} x^{-2/3}+O(x^{-1/3}). \end{veqnarray} Note that these formulae are only valid if there is no other shock in $[y_1-x/2,y_1+x/2]$. Since the probability of having another shock in $[y_1-x/2,y_1+x/2]$ is at most $O(x)$, the errors we incur by using (\ref{eq:4.9.1})-(\ref{eq:4.9.2}) are of higher order. Recall that \begin{veqnarray} \nonumber S_1(s,\xi_+,\xi_-,x)= \int\frac{d\lambda d\mu_+d\mu_-}{(2\pi)^3} {\rm e}^{i\lambda s+i\mu_+\xi_++i\mu_-\xi_-}\Omega(\lambda,\mu_+,\mu_-,x), \end{veqnarray} where \begin{veqnarray} \nonumber &&\sigma_1\Omega(\lambda,\mu_+,\mu_-,x) \\ &=&\biggl\<\sum_k {\rm e}^{-i\lambda s(y_k,x,t_k)-i\mu_+\xi(y_k+x/2,t_k) -i\mu_-\xi(y_k-x/2,t_k)}\delta(z-y_k)\delta(t-t_k) \biggr\>. \end{veqnarray} $\Omega$ is the characteristic function associated with $S_1$. Similarly \begin{veqnarray} \nonumber \Omega(\bar\lambda x^{-1/3},\bar\mu_+x^{2/3},\bar\mu_-x^{2/3},x), \end{veqnarray} is the characteristic function associated with the rescaled PDF \begin{veqnarray} \nonumber x S_1(\bar sx^{1/3},\bar\xi_+x^{-2/3},\bar\xi_-x^{-2/3},x). \end{veqnarray} We evaluate $\Omega(\bar\lambda x^{-1/3},\bar\mu_+x^{2/3},\bar\mu_-x^{2/3},x)$ in the limit as $x\to0$ using (\ref{eq:4.9.1}) and (\ref{eq:4.9.2}) for $s(y_1,x,t_1)$, $\xi(y_1+x/2,t_1)$. This gives \begin{veqnarray} \nonumber &&\sigma_1\Omega(\bar\lambda x^{-1/3},\bar\mu_+x^{2/3},\bar\mu_-x^{2/3},x)\\ &=& \biggl\<\sum_k{\rm e}^{-i\bar\lambda b - i(\bar\mu_++\bar\mu_-)b/3} \delta(z-y_k)\delta(t-t_k)\biggr\>+O(x^{1/3}). \end{veqnarray} In the limit as $x\to0$, $b$, $(y_k,t_k)$ are the only random quantities to be averaged over. Furthermore $b$ is statistically independent of $(y_k,t_k)$ because of statistical homogeneity and stationarity. Let $P(b)$ be the PDF of $b$. Then \begin{veqnarray} \nonumber &&\sigma_1\lim_{x\to0}\Omega(\bar\lambda x^{-1/3},\bar\mu_+x^{2/3},\bar\mu_-x^{2/3},x)\\ &=& \biggl\<\sum_k \delta(z-y_k)\delta(t-t_k)\biggr\> \int_{-\infty}^0\!\! db \ P(b){\rm e}^{-i\bar\lambda b -i(\bar\mu_++\bar\mu_-)b/3}\\ &=& \sigma_1 \int_{-\infty}^0\!\! db \ P(b){\rm e}^{-i\bar\lambda b -i(\bar\mu_++\bar\mu_-)b/3}. \end{veqnarray} Direct evaluation of this expression gives (\ref{eq:4.19}) in the variables $(\bar s,\bar\xi_+,\bar\xi_-)$. \enddemo \subsection{Connection with the geometric picture} \label{sec:4.6} Here we compute directly the contribution to $F$ in the neighborhood of shock creation. This is a reformulation of the argument presented in \cite{ekhma97} in terms of quantities defined in the present paper. Assume a shock is created at time $t=0$, position $x=y_1$, and with velocity $u=u_1$. Then locally (compare (\ref{eq:3.31.0})) \begin{veqnarray} \nonumber x=y_1+(u-u_1)t+a(u-u_1)^3+O((u-u_1)^2t), \end{veqnarray} where $a\le0$ is a random quantity. For the purpose of comparison with (\ref{eq:4.19.b}), it is useful to set $a=(2/b)^3$. Since for $t\ll1$ to leading order the shock is located at $x=y_1$, to leading order $u_-(y_1,t)$, $u_+(y_1,t)$ are solution of $0=(u-u_1)t-(2(u-u_1)/b)^3$. Thus \begin{veqnarray} \nonumber &&u_\pm(y_1,t)=u_1\mp \Bigl(\frac{|b|^3t}{8}\Bigr)^{1/2}+O(t), \end{veqnarray} \begin{veqnarray} \label{eq:4.31} &&s(y_1,t) = -\Bigl(\frac{|b|^3t}{2}\Bigr)^{1/2}+O(t) \end{veqnarray} Similarly, to leading order $\xi_-(y_1,t)$, $\xi_+(y_1,t)$ are solution of $1=\xi t-3(2/b)^3(u_\pm-u_1)^2\xi$. Thus \begin{veqnarray} \label{eq:4.32} \xi_\pm (y_1,t) = -\frac{1}{2t}+O(1). \end{veqnarray} Recall that from (\ref{eq:4.9.l2}) (using $\xi_+(x,t)\stackrel{\hbox{\tiny law}}{=} \xi_-(-x,t)$) \begin{veqnarray} \nonumber F(\xi)&=&\int\frac{d\mu}{2\pi} {\rm e}^{i\mu\xi}\biggl\<\sum_j s(z,t) {\rm e}^{-i\mu\xi_+(z,t)}\delta(z-y_j) \biggr\>\\ &=&\biggl\<\sum_j s(z,t) \xi-\xi_+(z,t)\delta(z-y_j) \biggr\>, \end{veqnarray} Under the assumption of ergodicity with respect to time-translation, $F(\xi)$ can be evaluated from \begin{veqnarray} \nonumber F(\xi)&=&\lim_{L,T\to+\infty}\frac{1}{2LT} \int_0^T\!\!dt \int_{-L}^{L}dz \sum_{j} s(z,t) \xi-\xi_+(z,t)\delta(z-y_j)\\ &=&\lim_{L,T\to+\infty}\frac{1}{2LT} \int_0^T\!\!dt \sum_{j=1}^{N(t)} s(y_j,t) \xi-\xi_+(y_j,t), \end{veqnarray} where $N(t)$ is number of shocks in $[-L,L]$ at time $t$. The contribution to $F$ near shock creation points, say $F_1$, can be evaluated for large negative $\xi$ using (\ref{eq:4.31}), (\ref{eq:4.32}) for $s(y_1,t)$, $\xi_+(y_1,t)$. This gives in the limit as $\xi\to-\infty$ \begin{veqnarray} \nonumber F_1(\xi)\sim -\sigma_1\int_{-\infty}^0\!\!db P(b)\int_0^{+\infty} \!\! dt \ \Bigl(\frac{|b|^3t}{2}\Bigr)^{1/2} \delta \left(\xi+\frac{1}{2t}\right)=-C |\xi|^{-5/2}, \end{veqnarray} where $C = 2^{-1/2} \sigma_1\int_{-\infty}^0 db \, |b|^{3/2} P(b)$. Comparing with (\ref{eq:4.19.b}), we conclude that $F_1(\xi)= F(\xi)$ to leading order. \section{Conclusions} \label{sec:5} To recapitulate the highlights of this paper, by writing down and working with the master equations in the inviscid limit, we have shown that the scaling of the structure functions is related to the shocks which are the singular structure in the limiting flow. The scaling of the PDFs, on the other hand, is related to the shock creation and collision points, which are singularities on the singular structures. The present paper provides a framework within which various statistical quantities of the stochastic Burgers equation can be calculated using self-consistent asymptotics without making closure assumptions. The main examples used here are the asymptotic behavior of structure functions and the PDF of the velocity gradient. It seems likely that other statistical quantities, such as the tails of the velocity PDF and the PDF for velocity difference, can also be analyzed in the present framework by exploiting further the source terms in (\ref{eq:3.6}), (\ref{eq:3.20}).
\section{Introduction.} Simple dynamical DNA models have been recently studied \cite{Gaeta} to understand the possible basic mechanisms of some biological processes such as denaturation, initiation of the transcription, and transcription. The thermal denaturation has been investigated by Peyrard, Bishop and co-workers \cite{Peyrard1,Peyrard5} with a unidimensional model (PB model) which allows local openings of the hydrogen bonds and formation of denaturation bubbles. The local openings can be analytically described as breather-like objects of small amplitude, which have nevertheless interesting properties: as long as their amplitude is small enough, they can move along the chain, collect energy and grow \cite{Peyrard7}. They can also be trapped by some local dishomogeneities \cite{Peyrard6}, which suggests that the properties of breathers could allow the formation of the transcription bubble after the interaction with the bound RNA-polymerase. However, in a realistic description of DNA dynamics, the topological constraints related to the helicoidal structure of the molecule cannot be ignored. Activation or repression effects caused by conformational changes and, in particular, prevention of transcription due to a large positive excess in twist, are known and largely investigated in biology. In fact, during all processes in which the DNA base-pairs open, a local unwinding of the helix follows for topological reasons. Consequently a local extra-twist accumulates at the two ends of the bubble and induces a long range elastic stress. Although the mechanical properties of DNA have recently been the object of a renewed interest\cite{Marko1,Marko2}, previous attempts to take into account the helicoidal structure in the models have been limited to the introduction of the forces that can appear due to the proximity in space of bases which are non-adjacent in the sequence \cite{Gaeta}. Our aim here is to build a simple model that takes into account the twist-opening interactions due to the helicoidal molecular geometry. Such a model provides an extension of the PB approach towards a more realistic description of biological processes. It can also be useful for more general studies of the interaction between geometrical conformation and dynamical properties of the molecule referring to the recent mechanical experiments on DNA \cite{SmithCui,StrickAllemand,ens}. In designing such a model each base will be considered as a single, non-deformable object. We introduce two degrees of freedom per base-pair: one radial variable related to the motion of the bases along the diameter that joins their attachment points to the helicoidal backbone, and a twist angle of each base-pair defined by the angle between this diameter and a reference direction. This twist angle, which increases from one base-pair to the next one, is responsible for the helicoidal structure of the molecule. From the resulting Lagrangian the equation of motion for small displacements with respect to the equilibrium position are written. We then derive analytical approximated solutions for the small amplitude nonlinear distortions of the molecule. As suggested by the helicoidal geometry we show that the radial breather-like opening is associated to an angular un-twisting of the molecule. We finally present the results of numerical simulations of the dynamics showing that these solutions are stable for long periods of time and can move along the chain. The analytical method of derivation of the envelope soliton solution of the equation of motion used in this work has been proposed and developed extensively in a previous paper \cite{noi}. The reader is referred to this work to better follow the technical steps of our calculation. \section{The helicoidal model.} In the PB model \cite{Peyrard1}, the bases are point masses allowed to move only in the direction of the hydrogen bonds that connect them. A Morse potential describes the effect of these bonds, while neighboring bases along the same strand are harmonically coupled, simulating the stacking interactions. Similarly, in our model the group made of a sugar ring and its connected base is treated simply as a point mass (without distinction between the different base types); the phosphate backbone between two base-pairs is modelised as an elastic rod. The additional twist motion is now introduced by allowing the two bases in each pair to move in the base-pair plane instead of constraining them on a line. It is convenient to choose a polar coordinate system (Fig. 1). The model does not attempt to describe the acoustic motions of the molecule since only the stretching of the base-pair distance is considered. This amounts to fixing the center of mass of the base pair, i.e. the two bases in a pair are constrained to move symmetrically with respect to the axis of the molecule. Then, to describe the stretching of a base pair and the variation of the helicoidal twist we need only two degrees of freedom per base pair: the coordinates ${r}_{n}$ and ${\varphi}_{n}$ of one of the two bases with respect to a fixed reference frame. As in the PB model, a Morse potential describes the hydrogen bonds linking bases in a pair with an equilibrium distance $R_0$. A proper choice of the coupling between radial and angular variables has to reproduce the equilibrium helicoidal structure. In DNA, the latter originates from the competition between the hydrophobic effect (that tends to eliminate water from the core of the molecule by bringing the neighboring base-pair planes closer) the electrostatic repulsion between neighboring base planes (which has the opposite effect) and the rigidity of the two strands (that separates the external ends of the base pairs by essentially a fixed length related to the phosphate length) \cite{Calladine}. We combine the first two forces in a unique stacking effect that fixes base-plane distance $h$ in the model. As the equilibrium backbone length $L$ is greater than $h$, it is then necessary to incline the strands in the typical helicoidal structure to minimize the energy. The geometrical parameters $h$ and $R_0$ have been chosen to match the structure of B-DNA. Then, in order to impose an equilibrium twist angle $ {\varphi}_{n} -{\varphi}_{n-1} =\pm {\Theta}_{0}$ between two consecutive base pairs, we select the equilibrium length $L$ of the springs representing the phosphate backbone to be \begin{eqnarray} \label{L} = \sqrt{{h}^{2}+ 4{R}_{0}^{2}\sin^2\big({{\Theta}_{0}\over 2}\big) }\, >\, h \end{eqnarray} \smallskip However, because of the invariance of $L$ with respect to the sign of ${\Theta}_{0}$, the elastic rods rigidity is not sufficient to guarantee the correct helicoidal shape. A {\it zig-zag} structure with a random succession of $\pm {\Theta}_{0}$ would as well minimize the energy of the system. This can be avoided by adding a three-body curvature term to the Lagrangian, that imposes a continuity between the differences ${\varphi}_{n} -{\varphi}_{n-1}$ and ${\varphi}_{n+1} -{\varphi}_{n}$. The final Lagrangian is written considering the expression of the three-dimensional distance between the two bases along the strand (Fig. 1); it reads \begin{eqnarray} \label{HBC} {\cal L} & = & \sum_{n} \big( m{\dot{{r}_{n}}}^{2}+m{{r}_{n}}^{2}{\dot{{\varphi}_{n}}}^{2}\big) - D \big( e^{-\alpha({r}_{n} - {R}_{0})} -1 \big)^2 \nonumber \\ &-& \sum_{n} K {\big( \sqrt{{h}^2 + {r}_{n-1}^2 + {r}_{n}^2 - 2 {r}_{n-1} {r}_{n} \cos ({\varphi}_{n} - {\varphi}_{n-1})} - L \big)}^2 \nonumber \\ &-& \sum_{n} G_0\,{\big( {\varphi}_{n+1}+{\varphi}_{n-1}-2{\varphi}_{n} \big)}^{2} \; \end{eqnarray} where $m$ is the base mass, $D$ and $\alpha$ are the depth and width of the Morse potential well (without distinction between double and triple hydrogen bonds), $K$ is the backbone elastic constant and $G_0$ the backbone curvature constant. The results presented below have been obtained with, $m=300 u.m.a.$, $D=0.04 eV$, $\alpha=4.45$ \AA$^{-1}$ that are the same valued adopted in the PB model, and with $K=1.0 eV $ \AA$^{-2}$ and $G_0=K R_0^2 / 2$. We intend to refine these values in future statistical mechanics studies by comparing the predictions of the present model with some available experimental data, {\em eg.} the temperature of DNA denaturation and the rigidity of the molecule \cite{Mezard}. For the geometrical parameters we adopt the B-DNA values $R_0 \approx 10.0$ \AA, ${\Theta}_{0}=36^{\circ} $ and $h=3.4$ \AA. One can simplify the Lagrangian by introducing the adimensional variables $r'_n=\alpha r_n$, the rescaled time $t'=\sqrt{D \alpha^2/m}\, t$ and the renormalized parameters \begin{eqnarray} K'&=&K/D \alpha^2 \nonumber \\ {\cal G'}&=&G_0/D \nonumber \\ R'_0&=&\alpha R_0 \nonumber \\ h'&=&\alpha h \nonumber \\ L' &=& \alpha L\,. \end{eqnarray} In the following the new variables will be written without primes. We perform an expansion of the energy of the backbone springs in (\ref{HBC}) up to the second order around the equilibrium position: \begin{eqnarray} y_n &=& r_n-R_0\,\\ {\phi}_{n} &=&R_0 (\,{\varphi}_{n}-n{\Theta}_{0}\,)\,, \end{eqnarray} \smallskip obtaining \begin{eqnarray} \label{appspring} {\big( \sqrt{{h}^2 + {r}_{n-1}^2 + {r}_{n}^2 - 2 {r}_{n-1} {r}_{n} \cos ({\varphi}_{n} - {\varphi}_{n-1})} -L\big)^2} \simeq \nonumber \\ {{ R_0^2 }\over{L^2}} {\big[ \sin{{\Theta}_{0}} ({\phi}_{n}-{\phi}_{n-1}) +(y_n+y_{n-1})(1-\cos{\Theta_0}) \big]}^{2} \,. \end{eqnarray} \smallskip We also expand the Morse potential up to the fourth order in $y_n$ \begin{equation} \big( e^{- {y}_{n}} -1 \big)^2 = \frac{1}{2} y_n^2- \frac{1}{2} y_n^3 +\frac{7}{24} y_n^4 + O({y_n}^5)\,. \end{equation} The difference in the order of the two expansions is consistent with our parameter choice. With these expansions, the equations of motion become: \begin{eqnarray} \label{eqm1} {\ddot y}_n &=&\big(1+{{y_n}\over{R_0}}\big){{1}\over{R_0}}{\dot{\phi}}_{n}^{2} - (y_n- {{3}\over{2}} y_n^2 +{{7}\over{6}}y_n^3 ) -K_{yy} ({y}_{n+1} + {y}_{n-1} + 2 {y}_{n}) \nonumber \\ &&-{{K_{y\phi}}\over{2}}({\phi}_{n+1} -{\phi}_{n-1}) \\ \label{eqm2} {\ddot{\phi}}_{n} &=& K_{\phi \phi} ({\phi}_{n+1}+{\phi}_{n-1}-2{\phi}_{n}) \nonumber \\ &-& G ( {\phi}_{n+2}+{\phi}_{n-2}-4{\phi}_{n+1}-4{\phi}_{n-1}+6{\phi}_{n}) \nonumber \\ &&+{{K_{y\phi}}\over{2}}({y}_{n+1} -{y}_{n-1})- \frac{2}{R_0}{\dot{y}}_{n}{\dot{\phi}}_{n}- \frac{2}{R_0}{{y}}_{n}{\ddot{\phi}}_{n}- \frac{2}{R_0^2}y_n {\dot{y}}_{n}{\dot{\phi}}_{n} -\frac{1}{R_0^2}y_n^2 {\ddot{\phi}}_{n} \end{eqnarray} where \begin{eqnarray} K_{yy} &=& \big( K R_0^2 / L^2 \big){(1 - \cos{\Theta}_{0})}^{2}\;,\\ K_{\phi\phi} &=& \big( K R_0^2 / L^2 \big){(\sin^2{\Theta}_{0})}\;, \\ K_{y\phi}&=&2 \big(K R_0^2 / L^2 \big)(\sin{\Theta}_{0})(1-\cos{\Theta}_{0})\; \\ G &=& {\cal G} / R_0^2 \, . \end{eqnarray} The constants $K_{yy},K_{\phi \phi}$ are the effective elastic constants respectively for the base pairs opening and the twist rotation, $K_{y\phi}$ is the coupling constant between stretching and twist; these three constant are geometrically related. \section{Localized breather-like solutions.} We now look for nonlinear localized solutions, characterized by the propagation of a coherent collective structure on a time scale greater than the time scale of the vibrations of each particle around its equilibrium position. According to the method developped in \cite{noi} we first solve for a wave packet solution of the linearized system with weak dispersion. This amounts to solving \begin{equation} \label{d} (\hat{J}(q)- \omega^{2}_{l}(q))\vec{V}_{l}(q) =0 \end{equation} where $l=+,-$ is the branch index and \begin{equation} \hat{J}(q)= \left(\matrix{1+2K_{yy}(1+\cos{q}) & iK_{y \phi} \sin{q} \cr -iK_{y \phi} \sin{q} & 2 K_{\phi \phi}(1-cos{q}) + G( 6-8\cos{q}+2\cos{2q})}\right) \equiv \left(\matrix{a & c \cr c^* & b}\right) \label{reldisp} \end{equation} finding the eigenvalues and eigenvectors \begin{equation} \omega^2_{\pm}(q)=\frac{1}{2}(a+b\pm\sqrt{(a-b)^2+4|c|^2}\;) \end{equation} \begin{equation} \label{autovettori1} \vec{V}^{\pm}(q)= {\cal N_{\pm}} \left( \barr{c} 1 \\ \frac{a-\omega_{\pm}^2} {-c} \end{array} \right) \end{equation} where ${\cal N_{\pm} }$ are the vector norms. Then we apply a perturbative expansion of the system (\ref{d}) around one normal mode. We have chosen as carrier wave a mode ($q_0,\omega_{+}=\omega_{+}(q_0)$) on the optical branch and with a small wave number that corresponds to a prevalently radial excitation weakly oscillating. In fact, keeping in mind DNA opening, we shall calculate the behaviour of a localized bubble-like radial distorsion and the induced angular dynamics. We obtain the wave-packet group velocity from the first order eigenvalue correction \begin{equation} \omega_{+}^{(1)}= \frac{ {\vec{V}^{+*}} \hat{J'} \vec{V}^{+}}{2 \omega_{+}}= \frac{1}{2 \omega_{+}}(a' |{V}^{+}_1|^2+c'{V}^{+*}_1{V}^{+}_2+ c'^{*}{V}^{+*}_2{V}^{+}_1+b'|{V}^{+}_2|^2) \quad , \end{equation} where the primes indicate the derivatives with respect to $q$ calculated at $q_0$. We then obtain the correction to the eigenvector $\vec{V}^{+}$ \begin{equation} \vec{V}^{(1)}=\alpha \vec{V}^{-} \end{equation} \begin{equation} \alpha= \frac{ {\vec{V}^{-*}} \hat{J}' \vec{V}^{+}} {\omega_{+}^2-\omega_{-}^2}= \frac{a' {V}^{-*}_1 {V}^{+}_1 +c'{V}^{-*}_1{V}^{+}_2+ c'^*{V}^{-*}_2{V}^{+}_1+b'{V}^{-*}_2 {V}^{+}_2 }{\omega_{+}^2-\omega_{-}^2} \end{equation} and from the second order eigenvalue correction the group velocity dispersion is obtained as \begin{eqnarray} \label{os} \omega_{+}^{(2)}&=& \frac{1}{\omega_{+}} ({\vec{V}}^{+*} \frac{\hat{J}''}{2} \vec{V}^{+} -{\omega_{+}^{(1)}}^2+ \frac{| {\vec{V}^{-*}} \hat{J}' \vec{V}^{+}|^2} { \omega^2_{+} -\omega^2_{-}})\ \nonumber \\ &=& \frac {1}{\omega_{+}}(\frac{1}{2}(a'' |{V}^{+}_1|^2+c''{V}^{+*}_1{V}^{+}_2+ c''^*{V}^{+*}_2{V}^{+}_1+b''|{V}^{+}_2|^2)- {\omega_{+}^{(1)}}^2 +|\alpha|^2 (\omega_{+}^2-\omega_{-}^2)) \end{eqnarray} We now take into account the nonlinearity. We look for a small amplitude solution. The expansion parameter $\epsilon$ for the solution is introduced to solve the equation of motion at increasing order of accuracy inserting the nonlinear terms in a progressive way. This iterative expansion is combined with the expansion in multiple scales for the weak dispersive wave packet. The latter requires the introduction of the variables $x_1=\epsilon x, t_1= \epsilon t, t_2=\epsilon^2 t$ for the slowly varying amplitudes, where $\epsilon$ is the same parameter as before. We look then for a solution of the form \begin{eqnarray} \label{wpc} \vec{E}(n)&=&\epsilon e^{i(q_0n_0-\omega_{+}t_0)} (\vec{V}^{+}-i \epsilon \vec {V}^{(1)} \frac{\partial }{\partial x_1}) A(x_1,t_1,t_2)+\epsilon \vec\sigma(x_1,t_1,t_2) \nonumber \\ &+&\epsilon^2 e^{2i(q_0n_0-\omega_{+}t_0)} \vec\gamma(x_1,t_1,t_2) +\epsilon^2 \vec\mu(x_1,t_1,t_2) \, \end{eqnarray} where the first term is the wave packet, the second term arise because the linear system admits a constant solution with a non zero second component, corresponding to the null column of $\hat{J}(0)$. The constant and second harmonic $\epsilon^{2}$ terms are induced by the quadratic nonlinearity in the equation of motion. We have now to determine the slowly variyng amplitudes $ A(x_1,t_1,t_2),\vec\sigma(x_1,t_1,t_2), \vec\gamma(x_1,t_1,t_2),\vec\mu(x_1,t_1,t_2)$ From the $O(\epsilon^2)$ equation of motion we obtain for the second harmonic terms the system of equations: \begin{equation} (\hat{J}(2q_0)- 4\omega_{+}^2)\vec{\gamma}= \left(\matrix{3/2 {V_1^{+}}^2 - \frac{\omega_{+}^2}{R_0}{V_2^{+}}^2 \cr \frac{4\omega_{+}^2}{R_0}V_1^{+} V_2^{+}} \right) \;A^2 \; . \end{equation} Its solution may be written as $\vec{\gamma}=\vec{\gamma_c} A^2$ and for the constant terms \begin{equation} \hat{J}(0)\vec\mu -i\hat{J'}(0)\frac{\partial}{\partial x_1} \vec\sigma =\left(\matrix{3 |{V_1^{+}}|^2 + \frac{2\omega_{+}^2}{R_0}|{V_2^{+}}|^2 \cr 0} \right)\;|A|^2 \quad . \end{equation} Inserting the value of $\hat J (0)$ from (\ref{reldisp}) in the above matricial equation, we obtain the identity \begin{equation} \label{mu1} (1+4K_{yy})\mu_1+K_{y\phi}\frac{\partial}{\partial x_1}\sigma_2= (3 |{V_1^{+}}|^2 + \frac{2\omega_{+}^2}{R_0}|{V_2^{+}}|^2) \;|A|^2 \end{equation} that does not allow to determine the two unknown variables. We have thus to consider the $O(\epsilon^3)$ system of equations for the constant terms \begin{equation} -i\hat{J'}(0) \frac{\partial}{\partial x_1}\vec\mu- \hat{J''}(0) \frac{\partial^2}{\partial x_1^2} \vec\sigma= \left(\matrix{ \frac{2i\omega_{+}}{R_0}|{V_2^{+}}|^2( A^* \frac{\partial A}{\partial t_1}-\frac{\partial A^*}{\partial t_1}A) \cr \frac{-2i\omega_{+}}{R_0} ({V_2^{+}}^* V_1^{+}-{V_1^{+}}^* V_2^{+}) \frac{\partial}{\partial t_1} |A|^2}\right) \end{equation} corresponding to the two equations: \begin{eqnarray} \label{mu2} K_{y \phi} \frac{\partial}{\partial x_1}\mu_2 &=& \frac{2i\omega_{+}}{R_0}|{V_2^{+}}|^2( A^* \frac{\partial A}{\partial t_1}-\frac{\partial A^*}{\partial t_1}A) \\ \label{sigma2} -K_{y\phi}\frac{\partial}{\partial x_1}\mu_1- 2K_{\phi \phi} \frac{\partial^2}{\partial x_1^2} \sigma_2 &=& \frac{-2i\omega_{+}}{R_0} ({V_2^{+}}^* V_1^{+}-{V_1^{+}}^* V_2^{+}) \frac{\partial}{\partial t_1} |A|^2 \end{eqnarray} The equation (\ref{mu2}) gives $\mu_2$ as a function of $A$. The equation (\ref{sigma2}) can be integrated using the wave packet property $\frac{\partial A}{\partial t_1}= - \omega_{+}^{(1)}\frac{\partial A}{\partial x_1}$. We then obtain from (\ref{mu1}), (\ref{sigma2}) the following system for $\mu_1$ and $\sigma_2$, \begin{equation} \left(\matrix{ 1+4K_{yy} &K_{y \phi} \cr -K_{y \phi} &-2K_{\phi\phi}}\right)\; \left(\matrix{\mu_1 \cr \frac{\partial \sigma_1}{\partial x_1}}\right)= \left(\matrix{3 |{V_1^{+}}|^2 + \frac{2\omega_{+}^2}{R_0}|{V_2^{+}}|^2 \cr \frac{2i\omega_{+}\omega_{+}^{(1)} }{R_0} ({V_2^{+}}^* V_1^{+}-{V_1^{+}}^* V_2^{+})}\right)\;|A|^2 \quad , \end{equation} the solutions of which may be written as \begin{eqnarray} \mu_1={\mu_1}_c |A|^2 \\ \sigma_2=\sigma_c\int|A|^2 dx_1 \end{eqnarray} The nonlinear $O(\epsilon^3)$ terms in $\exp{i(q_0n_0-\omega_{+}t_0)}$ in the equation of motion give rise to the nonlinearity that balances the group velocity dispersion of the wave packet. We indeed obtain the following Non Linear Schr\"odinger (NLS) equation for the envelope $A$ expressed in a frame moving at velocity $\omega_{+}^{(1)}$ (variables $S = x_1-\omega_{+}^{(1)} t_1,\tau=t_2 $): \begin{equation} \label{NLScr} i A_{\tau} + P A_{SS} + Q {|A|}^2 A = 0 \end{equation} with \begin{eqnarray} P &=& \frac{\omega_{+}^{(2)}}{2} \;, \\ Q &=& {{{ (Q_1 {V_{1}^+}^* + Q_2 {V_{2}^+}^*) }} \over { 2 \omega_+ } }\;, \end{eqnarray} where \begin{eqnarray} Q_1 &=& -\frac{7}{2}|V_{1}^+|^2 V_{1}^+ +\frac{\omega_{+}^{2}}{R_0^2} (2 V_{1}^+ |V_{2}^+|^2 -{V_{1}^+}^* {V_{2}^+}^2)+\nonumber\\ &&3{V_{1}^+}^* {\gamma_1}_c+\frac{4\omega_{+}^{2}}{R_0}{V_{2}^+}^*{\gamma_2}_c +3 V_{1}^+{\mu_1}_c \\ Q_2 &=& \frac{\omega_{+}^{2}}{R_0}(4{V_{1}^+}^* {\gamma_2}_c- 2{V_{2}^+}^* {\gamma_1}_c)+ \frac{2\omega_{+}^{2}}{R_0} V_{2}^+{\mu_1}_c - \nonumber\\ && \frac{2\omega_{+}^{2}}{R_0^2} {V_{1}^+}^2 {V_{2}^+}^* + \frac{\omega_{+}^{2}}{R_0^2}(2|V_{1}^+|^2 V_{2}^+ +{V_{1}^+}^2 {V_{2}^+}^*) \quad . \end{eqnarray} If $PQ > 0$, equation (\ref{NLScr}) has an envelope soliton solution \begin{equation} \label{F} A(S,\tau) = {\cal A}\ \mbox{sech}{ [ {{1}\over{L_e}} (S-u_e \tau)]} \exp{ [ i {{u_e}\over{2P}} (S-u_c \tau)]} \end{equation} where \begin{eqnarray} {\cal A} = \sqrt{ {{u_e^2 - 2 u_e u_c}\over{2PQ}} } \\ L_e = {{2P}\over{\sqrt{u_e^2 - 2 u_e u_c}}} \end{eqnarray} are respectively the amplitude and the width of the curve. Once the NLS equation is solved for $A(S, \tau )$, we have $\sigma_2$, $\vec{\gamma}$, $\vec\mu$ (solving (\ref{mu2}) we obtain $\mu_2={\mu_2}_c \int |A|^2 dx_1$) and then the complete solution \begin{eqnarray} y &=& \epsilon (V_1^+ -i\epsilon V_1^{(1)} \frac{\partial} {\partial x_1}) A e^{i \theta} + c.c. +\epsilon^2 {\gamma_1}_c A^2 e^{2 i \theta} + c.c. +\epsilon^2 {\mu_1}_c {|A|}^2 +O(\epsilon^3) \\ \phi &=& \epsilon \sigma_c \int {|A|}^{2} dx_1 + \epsilon ( V_2^+ -i\epsilon V_2^{(1)} \frac{\partial} {\partial x_1}) A e^{i \theta} + c.c. +\epsilon^2 {\gamma_2}_c A^2 e^{2 i \theta} + c.c.+ \epsilon^2 {\mu_2}_c \int {|A|}^{2} dx_1 +O(\epsilon^3) \, . \end{eqnarray} where $\theta= (q_0n_0-\omega_{+}t_0)\,.$ The final result can be rewritten in the following form: \begin{eqnarray} \label{y} y &=& 2\epsilon {V_1}^+ {\cal A} \mbox{sech}{ [ \eta (x-V_e t)]} \cos{( {\cal K}x - {\Omega}t )} \nonumber \\ &+& \epsilon^2 {V_1}^{(1)} {\cal A} \mbox{sech}{ [ \eta (x-V_e t)]} \left[\frac{-2}{ L_e}\tanh{(\eta(x - V_e t))} \sin{( {\cal K}x - {\Omega}t )} +{{u_e }\over{P}}\cos{( {\cal K}x - {\Omega}t )}\right] \nonumber \\ &+& 2\epsilon^2 {\gamma_1}_c {\cal A}^2 {\mbox{sech}}^2 { [ \eta (x-V_e t)]} \cos{(2 {\cal K}x - 2 {\Omega}t )} \nonumber \\ &+& {\mu_1}_c \epsilon^2 {\cal A}^2 {\mbox{sech}}^2 { [ \eta (x-V_e t)]} +O(\epsilon^3) \end{eqnarray} \begin{eqnarray} \label{phi} \phi &=& \epsilon (\sigma_c+\epsilon {\mu_2}_c) L_e {\cal A}^2 \tanh{(\eta(x - V_e t))} -2\epsilon {\cal A}|V_2^+| \mbox{sech}{ [ \eta (x-V_e t)]} \sin{( {\cal K}x - {\Omega}t )} \nonumber \\ &-& \epsilon^2 {V_2}^{(1)} {\cal A} \mbox{sech}{ [ \eta (x-V_e t)]} \left[\frac{2}{ L_e}\tanh{(\eta(x - V_e t))} \cos{( {\cal K}x - {\Omega}t )} +{{u_e }\over{P}}\sin{( {\cal K}x - {\Omega}t )}\right] \nonumber \\ &-& 2 \epsilon^2|{\gamma_2}_c| {\cal A}^2 {\mbox{sech}}^2 { [ \eta (x-V_e t)]} \big[ \sin{( 2{\cal K}x - 2{\Omega}t )} \big] +O(\epsilon^3) \end{eqnarray} where \begin{eqnarray} \eta &=& {{\epsilon}\over{L_e}} \\ V_e &=& {\omega_+}^{(1)} + \epsilon u_e \\ V_c &=&{\omega_+}^{(1)} + \epsilon u_c \\ {\cal K} &=& q_0 + {{\epsilon u_e}\over{2P}} \\ {\Omega} &=& \omega_+ + {{\epsilon u_e}\over{2P}} V_c \, . \end{eqnarray} Fig. 3 shows the analytical solution (\ref{y}),(\ref{phi}) in the original variables (not renormalized), as a function of $x$ and $t$. Fig. 4 displays the results of numerical simulations of the system described by the complete Lagrangian (\ref{L}) starting from the analytical solution (\ref{y}),(\ref{phi}) for $t=0$ as an initial condition. We have chosen the carrier wavevector $q_0=0.1$. The shape of the solution is best seen on Fig. 3a, 3b which show for a few periods the radial displacement $r_n(t)-R_0$ (Fig. 3a) and of the angular displacement $\varphi_n(t)$ (Fig. 3b), while Fig. 3c represents the twist angle, {\em i.e.} the difference between neighboring angles, $\Delta\varphi_n(t)$. As in the PB model the radial motion has the shape of an asymmetric breather, the stretching of the base-pair distance being larger than its compression, as expected from the asymmetry of the Morse potential. The corresponding angular motion clearly exhibits a kink structure due to the non oscillating term $\sigma (x,t)$; a small oscillating part due to the other terms in (\ref{phi}) is superimposed on it. The shape of the solution agrees with the geometrical properties of the molecule related to its helicoidal structure. It shows, as expected, that our model can indeed describe the local untwisting which should be coupled with the radial breather due to the geometrical constraints. To make this clearer we plot in Fig. 3c the evolution of the twist angle. In addition to the local opening, small periodic overtwists on the boundaries, coming from the oscillating part of the solutions are visible. The coupled breather and kink solutions move with a peak velocity $ V_e = \omega_{+}^{(1)} + \epsilon u_e$ without changing their internal structure. The long time evolution of the initial condition, calculated with the full Lagrangian and shown in Fig.~4, attests that the local excitation is very stable. It does not radiates energy with the exception of a small initial transient that could be eliminated with absorbing boundary conditions. During all the simulation time the total energy lost with the absorbing conditions is less than $ 10^{-5}$ of the total energy. It should be noticed that Fig. 4 does not show the actual oscillations due to the large sampling time used to produce the graph. The slow oscillation visible on Fig. 4 is due to a beating between the actual oscillation and the sampling time. Fig. 4 shows in fact about 2230 oscillations which proves the exceptional stability of the breather in the helicoidal structure and the ability of the geometrical nonlinearity and of the nonlinearity coming from the Morse potential to keep the energy localized. As our solution has been obtained with a small amplitude expansion, we should not expect it to be a good solution when its amplitude increases to much. This can be seen on the results of a numerical simulation starting from a much larger amplitude (Fig.~5). In this case the initial condition decays very fast and radiates energy before stabilizing to a breather which has still a pretty high amplitude and which is localized on about 6 sites (width at half height). Although it is very narrow the breather is still mobile in the system and it is accompanied by a sharp kink as expected from the geometrical constraints. The solution reached after the transient is stable as shown in Fig. 5c: although the simulation has been run with absorbing boundary conditions, the energy tends to stabilize once the radiated waves have been dissipated at the ends of the molecule. In all the solutions described above the breather frequency $\Omega$ remains in the gap between the two branches so that the breather does not emit acoustic phonons. \section{Conclusion} In conclusion we have proposed a two-variable helicoidal DNA model that, from a biological point of view, can be a better starting point than the PB model for the studies of all the DNA dynamical processes involving internal openings. Our model also describes the untwisting that accompanies the base-pair stretching because of the helicoidal molecular geometry. We analytically derive, using the method introduced in \cite{noi}, a breather like solution in the radial variable coupled with an angular solution dominated by a non oscillating kink-like term. A further step will be the study of the formation of this kind of excitation in presence of a thermal bath and of the onset of thermal denaturation. This will be useful in determining the model parameters by comparison with denaturation temperature experimental data. Furthermore we expect to find interesting thermodynamic properties due to long range effect generated by the helicoidal DNA structure. As far as biological processes are concerned, the model may however still lack some relevant features, in particular the possibility of bending and stretching of the molecule. In order to obtain a stretchable molecule axis we can simply replace the rigid constraint of fixed base planes distance $h$ with an elastic term which leads to the introduction of a third degree of freedom with a new elastic constant. This could be a link to some recent experimental studies of the mechanical response of a DNA molecule to some twisting and stretching forces \cite{SmithCui,StrickAllemand,ens}. Anyway the model contains in its possible dynamical behaviors some other important features, as {\em e.g.} the possibility of local transitions from right hand B-DNA to left hand Z-DNA, related to the curvature term in the Lagrangian depending on the constant $G_0$. Our model is, finally, a first attempt to insert in simple DNA dynamical model topological constraints that are of great interest in today's biological studies. \section*{Acknowledgments} We would like to thank Prof. A. Colosimo and Prof. S. Ruffo for helpful discussions. This work was initiated by discussions that took place at the Institute for Scientific Interchange (ISI) of Torino, Italy, which is gratefully acknowledged. \newpage
\section{Introduction} There is now strong evidence for atmospheric neutrino oscillations \cite{SK}, \cite{K} which confirms the earlier indications of the effect \cite{early}. The most recent analyses of Super-Kamiokande \cite{SK} involve the hypothesis of $\nu_{\mu}\rightarrow \nu_{\tau}$ oscillations with maximal mixing $\sin^2 2\theta_{23} =1$ and a mass splitting of $\Delta m_{23}^2 = 2.2\times 10^{-3}\ eV^2$. Using all their data sets analysed in different ways they quote $\sin^2 2\theta_{23} > 0.82$ and a mass splitting of $5\times 10^{-4}\ eV^2 <\Delta m_{23}^2 < 6\times 10^{-3}\ eV^2$ at 90\% confidence level. The evidence for solar neutrino oscillations is almost as strong \cite{SNU}. There are a panoply of experiments looking at different energy ranges, and the best fit to all of them has been narrowed down to two basic scenarios corresponding to either resonant oscillations $\nu_e \rightarrow \nu_0$ (where for example $\nu_0$ may be a linear combination of $\nu_{\mu} , \nu_{\tau}$) inside the Sun (MSW \cite{MSW}) or ``just-so'' oscillations in the vacuum between the Sun and the Earth \cite{justso1, justso2}. There are three MSW fits and one vacuum oscillation fit: (i) the small angle MSW solution is $\sin^2 2\theta_{12} \approx 5 \times 10^{-3}$ and $\Delta m_{12}^2 \approx 5\times 10^{-6}\ eV^2$; (ii) the large angle MSW solution is $\sin^2 2\theta_{12} \approx 0.76$ and $\Delta m_{12}^2 \approx 1.8\times 10^{-5}\ eV^2$; (iii) an additional MSW large angle solution exists with a lower probability \cite{BKS}; (iv) The vacuum oscillation solution is $\sin^2 2\theta_{12} \approx 0.75$ and $\Delta m_{12}^2 \approx 6.5\times 10^{-11}\ eV^2$ \cite{BKS}. The standard model has zero neutrino masses, so any indication of neutrino mass is very exciting since it represents new physics beyond the standard model. In this paper we shall assume the see-saw mechanism and no light sterile neutrinos. The see-saw mechanism \cite{seesaw} implies that the three light neutrino masses arise from some heavy ``right-handed neutrinos'' $N^p_R$ (in general there can be $Z$ gauge singlets with $p=1,\ldots Z$) with a $Z\times Z$ Majorana mass matrix $M^{pq}_{RR}$ whose entries take values at or below the unification scale $M_U \sim 10^{16}$ GeV. The presence of electroweak scale Dirac mass terms $m_{LR}^{ip}$ (a $3 \times Z$ matrix) connecting the left-handed neutrinos $\nu^i_L$ ($i=1,\ldots 3$) to the right-handed neutrinos $N^p_R$ then results in a very light see-saw suppressed effective $3\times 3$ Majorana mass matrix \begin{equation} m_{LL}=m_{LR}M_{RR}^{-1}m_{LR}^T \label{seesaw} \end{equation} for the left-handed neutrinos $\nu_L^i$, which are the light physical degrees of freedom observed by experiment. Not surprisingly, following the recent data, there has been a torrent of theoretical papers concerned with understanding how to extend the standard model in order to accomodate the atmospheric and solar neutrino data \cite{torrent}. Perhaps the minimal extension of the standard model capable of accounting for the atmospheric neutrino data involves the addition of a {\em single} right-handed neutrino $N_R$ \cite{SK1, SK2}. This is a special case of the general see-saw model with $Z=1$, so that $M_{RR}$ is a trivial $1 \times 1$ matrix and $m_{LR}$ is a $3 \times 1$ column matrix where $m_{LR}^T=(\lambda_{\nu_e}, \lambda_{\nu_{\mu}}, \lambda_{\nu_{\tau}})v_2$ with $v_2$ the vacuum expectation value of the Higgs field $H_2$ which is responsible for the neutrino Dirac masses, and the notation for the Yukawa couplings $\lambda_i$ indicates that we are in the charged lepton mass eigenstate basis $e_L, \mu_L, \tau_L$ with corresponding neutrinos $\nu_{e_L}, \nu_{\mu_L}, \nu_{\tau_L}$. Since $M_{RR}$ is trivially invertible the light effective mass matrix in Eq.\ref{seesaw} in the $\nu_{e_L}, \nu_{\mu_L}, \nu_{\tau_L}$ basis is simply given by \begin{equation} m_{LL}=\frac{m_{LR}m_{LR}^T}{M_{ RR}} = \left( \begin{array}{lll} \lambda_{\nu_e}^2 & \lambda_{\nu_e} \lambda_{\nu_{\mu}} & \lambda_{\nu_e} \lambda_{\nu_{\tau}} \\ \lambda_{\nu_e} \lambda_{\nu_{\mu}} & \lambda_{\nu_{\mu}}^2 & \lambda_{\nu_{\mu}} \lambda_{\nu_{\tau}} \\ \lambda_{\nu_e }\lambda_{\nu_{\tau}} & \lambda_{\nu_{\mu}} \lambda_{\nu_{\tau}} & \lambda_{\nu_{\tau}}^2 \end{array} \right)\frac{v_2^2}{M_{RR}}. \label{matrix} \end{equation} The matrix in Eq.\ref{matrix} has vanishing determinant which implies a zero eigenvalue. Furthermore the submatrix in the 23 sector has zero determinant which implies a second zero eigenvalue associated with this sector. In order to account for the Super-Kamiokande data we assumed \cite{SK1}: \begin{equation} \lambda_{\nu_e} \ll \lambda_{\nu_{\mu}} \approx \lambda_{\nu_{\tau}}. \label{hierarchy} \end{equation} In the $\lambda_{\nu_e}=0$ limit the matrix in Eq.\ref{matrix} has zeros along the first row and column, and so clearly $\nu_e$ is massless, and the other two eigenvectors are simply \begin{equation} \left( \begin{array}{l} \nu_0 \\ \nu_3 \end{array} \right) = \left( \begin{array}{ll} c_{23} & -s_{23}\\ s_{23} & c_{23} \end{array} \right) \left( \begin{array}{l} \nu_{\mu} \\ \nu_{\tau} \end{array} \right) \label{bbasis} \end{equation} where $t_{23}=\lambda_{\nu_{\mu}}/\lambda_{\nu_{\tau}}$, with $\nu_0$ being massless, due to the vanishing of the determinant of the 23 submatrix and $\nu_3$ having a mass $m_{\nu_3}=(\lambda_{\nu_{\mu}}^2 + \lambda_{\nu_{\tau}}^2) v_2^2/M_{RR}$. The Super-Kamiokande data is accounted for by choosing the parameters such that $t_{23} \sim 1$ and $m_{\nu_3} \sim 5\times 10^{-2}$ eV. In this approximation the atmospheric neutrino data is then consistent with $\nu_{\mu}\rightarrow \nu_{\tau}$ oscillations via two state mixing, between $\nu_3$ and $\nu_0$. Note how the single right-handed neutrino coupling to the 23 sector implies vanishing determinant of the 23 submatrix. This provides a natural explanation of both large 23 mixing angles and a hierarchy of neutrino masses in the 23 sector at the same time \cite{SK1}. In order to account for the solar neutrino data a small mass perturbation is required to lift the massless degeneracy of the two neutrinos $\nu_0 ,\nu_e$. In our original approach \cite{SK1} \footnote{ Another approach \cite{SK2} which does not rely on additional right-handed neutrinos is to use SUSY radiative corrections so that the one-loop corrected neutrino masses are not zero but of order $10^{-5}$ eV suitable for the vacuum oscillation solution.} we introduced additional right-handed neutrinos in order to provide a subdominant contribution to the effective mass matrix in Eq.\ref{matrix}. To be precise we assumed a single dominant right-handed neutrino below the unification scale, with additional right-handed neutrinos at the unification scale which lead to subdominant contributions to the effective neutrino mass matrix. By appealing to quark and lepton mass hierarchy we assumed that the additional subdominant right-handed neutrinos generate a contribution $m_{\nu_{\tau}}\approx m_t^2/M_U \approx 2\times 10^{-3}$ eV, where $m_t$ is the top quark mass. The effect of this is to give a mass perturbation to the 33 component of the mass matrix in Eq.\ref{matrix}, which results in $\nu_0 $ picking up a small mass, through its $\nu_{\tau}$ component, while $\nu_e $ remains massless. Solar neutrino oscillations then arise from $\nu_e \rightarrow \nu_0$ with the mass splitting in the right range for the small angle MSW solution, controlled by a small mixing angle $\theta_{12} \approx \lambda_{\nu_e}/\sqrt{\lambda_{\nu_{\mu}}^2 + \lambda_{\nu_{\tau}}^2}$. The main prediction of this scheme is of the neutrino oscillation $\nu_e \rightarrow \nu_3$ with a mass difference $\Delta m_{13}^2 \approx \Delta m_{23}^2$ determined by the Super-Kamiokande data and a mixing angle $\theta_{13} \approx \theta_{12}$ determined by the small angle MSW solution. Such oscillations may be observable at the proposed long baseline experiments via $\nu_3 \rightarrow \nu_e$ which implies $\nu_{\mu} \rightarrow \nu_e$ oscillations with $\sin^2 2\theta \approx 5 \times 10^{-3}$ (the small MSW angle) and $\Delta m^2 \approx 2.2\times 10^{-3}\ eV^2$ (the Super-Kamiokande square mass difference). It should be clear from the foregoing discussion that the motivation for single right-handed neutrino dominance (SRHND) is that the determinant of the 23 submatrix of Eq.\ref{matrix} approximately vanishes, leading to a natural explanation of {\em both} large neutrino mixing angles {\em and} hierarchical neutrino masses in the 23 sector {\em at the same time} \cite{SK1}. Although the explicit example of SRHND above was based on one of the right-handed neutrinos being lighter than the others, it is clear that the idea of SRHND is more general than this. In the present paper we shall define SRHND more generally as the requirement that a single right-handed neutrino gives the dominant contribution to the 23 submatrix of the light effective neutrino mass matrix. We shall propose SRHND as a general requirement and address the following two questions: \newline 1. What are the general conditions under which SRHND in the 23 block can arise and how can we quantify the contribution of the sub-dominant right-handed neutrinos which are responsible for breaking the massless degeneracy, and allowing the small angle MSW solution? \newline 2. How can we understand the pattern of neutrino Yukawa couplings in Eq.\ref{hierarchy} where the assumed equality $\lambda_{\nu_{\mu}} \approx \lambda_{\nu_{\tau}}$ is apparently at odds with the hierarchical Yukawa couplings in the quark and charged lepton sector? \newline In order to address the two questions above we shall discuss SRHND in the context of a $U(1)$ family symmetry. In fact neutrino masses and mixing angles have already been studied in the context of $U(1)$ family symmetry models but in the models that exist to date either SRHND is not present at all \cite{Ross}, \cite{Ramond}, or where it is present its presence has apparently gone unnoticed \cite{Altarelli}. \footnote{We should point out that the condition of the approximately vanishing subdeterminant was first clearly stated in ref.\cite{Altarelli}. However all the actual examples presented there correspond to a single right-handed neutrino giving the dominant contribution to the 23 block of the effective neutrino mass matrix, which is essentially the mechanism first proposed in ref.\cite{SK1}. Also note that SRHND has very recently been applied to an $SU(2)$ family symmetry model\cite{Barbieri}.} Where there is no SRHND, either the contribution to the 23 mixing angles coming from the neutrino matrix are small \cite{Ross}, or the 23 neutrino mass hierarchy is not described by the Wolfenstein expansion parameter \cite{Ramond}. Where the 23 neutrino mass hierarchies are described by the Wolfenstein expansion parameter {\em and} large 23 mixing angles naturally arise \cite{Altarelli}, we shall show that the physical reason why these models are successful is that a single right-handed neutrino is giving the dominant contribution to the 23 submatrix of $m_{LL}$. We shall give general conditions that theories with $U(1)$ family symmetry must satisfy in order to have SRHND and show that the models in \cite{Altarelli} satisfy these conditions. \section{MSSM with $Z$ Right-handed Neutrinos} To fix the notation, we assume the Yukawa terms of the minimal supersymmetric standard model (MSSM) augmented by $Z$ right-handed neutrinos, \begin{eqnarray} {\cal L}_{yuk}=\epsilon_{ab}\left[ -Y^u_{ij}H_u^aQ_i^bU^c_j +Y^d_{ij}H_d^aQ_i^bD^c_j +Y^e_{ij}H_d^aL_i^bE^c_j -Y^{\nu}_{ip}H_u^aL_i^bN^c_p + \frac{1}{2}M_{RR}^{pq}N^c_pN^c_q \right] \nonumber \\ +H.c. \label{MSSM} \end{eqnarray} where $\epsilon_{ab}=-\epsilon_{ba}$, $\epsilon_{12}=1$, and the remaining notation is standard except that the $Z$ right-handed neutrinos $N_R^p$ have been replaced by their CP conjugates $N^c_p$ with $p,q=1,\dots, Z$. When the two Higgs doublets get their vacuum expectation values (VEVS) $<H_u^2>=v_2$, $<H_d^1>=v_1$ with $\tan \beta \equiv v_2 /v_1$ we find the terms \begin{equation} {\cal L}_{yuk}=v_2Y^u_{ij}U_iU^c_j +v_1Y^d_{ij}D_iD^c_j +v_1Y^e_{ij}E_iE^c_j +v_2Y^{\nu}_{ip}N_iN^c_p + \frac{1}{2}M_{RR}^{pq}N^c_pN^c_q +H.c. \end{equation} Replacing CP conjugate fields we can write in a matrix notation \begin{equation} {\cal L}_{yuk}=\bar{U}_Lv_2Y^uU_R +\bar{D}_Lv_1Y^dD_R +\bar{E}_Lv_1Y^eE_R +\bar{N}_Lv_2Y^{\nu}N_R + \frac{1}{2}N^T_RM_{RR}N_R +H.c. \end{equation} where we have assumed that all the masses and Yukawa couplings are real and written $Y^\star =Y$. The diagonal mass matrices are given by the following unitary transformations \begin{eqnarray} v_2Y^u_{diag}=V_{uL}v_2Y^uV_{uR}^{\dag}={\rm diag (m_u,m_c,m_t)},\nonumber \\ v_1Y^d_{diag}=V_{dL}v_1Y^dV_{dR}^{\dag}={\rm diag(m_d,m_s,m_b)},\nonumber \\ v_1Y^e_{diag}=V_{eL}v_1Y^eV_{eR}^{\dag}={\rm diag(m_e,m_{\mu},m_{\tau})}, \nonumber \\ M_{RR}^{diag}=\Omega_{RR}M_{RR}\Omega_{RR}^{\dag} ={\rm diag(M_{R1},\ldots ,M_{RZ})}, \end{eqnarray} where the unitary transformations are also orthogonal. From Eq.\ref{seesaw} the light effective left-handed Majorana neutrino mass matrix is \begin{equation} m_{LL}=v_2^2Y_{\nu}M_{RR}^{-1}Y_{\nu}^T \label{mLL} \end{equation} Having constructed the light Majorana mass matrix it must then be diagonalised by unitary transformations, \begin{equation} m_{LL}^{diag}=V_{\nu L}m_{LL}V_{\nu L}^{\dag} ={\rm diag(m_{\nu_1},m_{\nu_2},m_{\nu_3})}. \end{equation} The CKM matrix is given by \begin{equation} V_{CKM}=V_{uL}V_{dL}^{\dag} \end{equation} and its leptonic analogue is the MNS matrix \cite{MNS} \begin{equation} V_{MNS}=V_{\nu L}V_{eL}^{\dag}. \end{equation} \section{Wolfenstein Expansions} The Wolfenstein parametrisation of the CKM matrix yields the approximate form \cite{Wolf}: \begin{equation} V_{CKM} \sim \left( \begin{array}{lll} 1 & \lambda & \lambda^3 \\ \lambda & 1 & \lambda^2 \\ \lambda^3 & \lambda^2 & 1 \end{array} \right) \label{Wolf} \end{equation} where $\lambda \approx V_{us} \approx 0.22$. The horizontal quark and lepton mass ratios may similarly be expanded in terms of the Wolfenstein parameter: \footnote{We follow the expansions in ref.\cite{Ramond} even though $\frac{m_e}{m_{\tau}} \sim \lambda^5$ is a better fit.} \begin{equation} \frac{m_u}{m_t} \sim \lambda^8, \ \ \frac{m_c}{m_t} \sim \lambda^4, \ \ \frac{m_d}{m_b} \sim \lambda^4, \ \ \frac{m_s}{m_b} \sim \lambda^2, \ \ \frac{m_e}{m_{\tau}} \sim \lambda^4, \ \ \frac{m_{\mu}}{m_{\tau}} \sim \lambda^2. \ \ \end{equation} Assuming the MSSM the vertical quark and lepton mass ratios at $M_U$ are \begin{equation} \frac{m_b}{m_t} \sim \lambda^3, \ \ \frac{m_b}{m_{\tau}} \sim 1. \ \ \end{equation} Assuming that $V_{CKM}\sim V_{uL} \sim V_{dL}$, and the diagonal elements of the Yukawa matrices are of the same order as the eigenvalues:\footnote{Again this is similar to ref.\cite{Ramond} except that we allow a more general $\tan \beta$ dependence} \begin{equation} Y^u \sim \left( \begin{array}{lll} \lambda^8 & \lambda^5 & \lambda^3 \\ - & \lambda^4 & \lambda^2 \\ - & - & 1 \end{array} \right), \ \ \ \ Y^d \sim \left( \begin{array}{lll} \lambda^4 & \lambda^3 & \lambda^3 \\ - & \lambda^2 & \lambda^2 \\ - & - & 1 \end{array} \right)\lambda^n \label{Ramond} \end{equation} where $\tan \beta \sim \lambda^{n-3}$. Note that the CKM matrix only gives information about the upper triangular parts of the quark Yukawa matrices. The MNS matrix is less well determined, but Super-Kamiokande tells us that $\theta_{23} \sim 1$ and the small angle MSW solution implies $\theta_{12} \sim \lambda^2$. In addition for $\Delta m^2>9 \times 10^{-4} \ eV^2$ (i.e. over most of the atmospheric range) CHOOZ \cite{CHOOZ} fails to observe $\nu_e \rightarrow \nu_3$ and excludes $\sin^2 2\theta_{13}>0.18$ or $\theta_{13}>0.22 $. Hence CHOOZ allows $\theta_{13}\leq \lambda^2 $. If we assume for the sake of argument that $\theta_{13}\sim \lambda^2 $ (recall that this is a prediction of SRHND which follows from Eqs.\ref{matrix} and \ref{hierarchy}) then $V_{MNS}$ is given by: \begin{equation} V_{MNS} \sim \left( \begin{array}{lll} 1 & \lambda^2 & \lambda^2 \\ \lambda^2 & 1 & 1 \\ \lambda^2 & 1 & 1 \end{array} \right) \label{Wolflep} \end{equation} Then, in a similar way to the quarks, assuming that $V_{MNS}\sim V_{\nu L} \sim V_{eL}$, and the diagonal elements of the charged lepton matrix are of the same order as the eigenvalues we deduce \begin{equation} Y^e \sim \left( \begin{array}{lll} \lambda^4 & \lambda^4 & \lambda^2 \\ - & \lambda^2 & 1 \\ - & - & 1 \end{array} \right) \lambda^n. \label{Ramond'} \end{equation} The same argument applied to $m_{LL}$ runs into trouble because the hierarchy between the second and third eigenvalues is apparently not consistent with $\theta_{23} \sim 1$. To be precise Super-Kamiokande tells us that $m_{\nu_3} \approx 5 \times 10^{-2}$ eV, and small angle MSW tells us that $m_{\nu_2} \approx 2 \times 10^{-3}$ eV, hence \begin{equation} \frac{m_{\nu_2}}{m_{\nu_3}}\sim \lambda^2. \label{23hierarchy} \end{equation} The problem is how to generate such a hierarchy in the presence of large neutrino mixing angles. Note that this problem can be avoided for the charged lepton matrix in Eq.\ref{Ramond'} due to the undetermined 32 element which can be small, but for the symmetric neutrino matrix it is a problem. Fortunately the solution is provided by SRHND which implies that $m_{LL}$ is given from Eqs.\ref{matrix} and \ref{hierarchy} as: \begin{equation} m_{LL} \sim \left( \begin{array}{lll} \lambda^4 & \lambda^{2} & \lambda^2 \\ \lambda^2 & 1 & 1 \\ \lambda^2 & 1 & 1 \end{array} \right) m_{\nu_3} \label{mLL''} \end{equation} It is clear that SRHND leads to the prediction \begin{equation} \frac{m_{\nu_1}}{m_{\nu_3}}\sim \lambda^4, \end{equation} in addition to the previously mentioned prediction $\theta_{13}\sim \lambda^2$. The key to obtaining the hierarchy in Eq.\ref{23hierarchy} from Eq.\ref{mLL''} is the requirement that the determinant of the 23 submatrix must vanish to order $\lambda^2$. Since this subdeterminant naturally vanishes for a single right-handed neutrino coupling to the 23 sector, as in Eq.\ref{matrix}, all that is required is for the subdominant right-handed neutrino to generate a perturbation to the masses in the 23 sector which are of order $\lambda^2$ smaller than the leading contribution. We shall now discuss how this can come about in the framework of theories with broken $U(1)$ family symmetry. \section{U(1) Family Symmetry} The idea of accounting for the fermion mass spectrum via a broken family symmetry has a long history \cite{FN}, \cite{textures}. For definiteness we shall focus on a particular class of model based on a single pseudo-anomalous $U(1)_X$ gauged family symmetry \cite{IR}. We assume that the $U(1)_X$ is broken by the equal VEVs of two MSSM singlets $\theta , \bar{\theta}$ which have vector-like charges $\pm 1$ \cite{IR}. Theories in which the $U(1)_X$ is broken by a chiral MSSM singlet $\chi $ which has charge of one sign only, say $+1$, have also been proposed \cite{shrock}, \cite{chiral}. In all these cases the $U(1)_X$ has anomalies in the effective low energy theory below $M_U$ but these are compensated by string theory effects at $M_U$ and the Green-Schwartz mechanism \cite{GS} provides a dimension-five interaction term, whose structure demands a specific pattern among the anomaly coefficients \cite{IR}: \begin{equation} A(SU(3)_c^2U(1)_X):A(SU(2)_L^2U(1)_X):A(U(1)_Y^2U(1)_X)=1:1:5/3 \label{anomalies} \end{equation} The $U(1)_X$ breaking scale is set by $<\theta >=<\bar{\theta} >$ where the VEVs arise from a Green-Schwartz computable Fayet-Illiopoulos $D$-term which determines these VEVs to be one or two orders of magnitude below $M_U$. Additional exotic matter which exists in vector-like pairs with opposite charges $\pm X_i$ at a heavy mass scale $M_V$ (generated by the VEVs of yet more singlets) allows the Wolfenstein parameter to be generated by the ratio \cite{IR} \begin{equation} \frac{<\theta >}{M_V}=\frac{<\bar{\theta} >}{M_V}= \lambda \approx 0.22 \label{expansion} \end{equation} The idea is that at tree-level the $U(1)_X$ family symmetry only permits third family Yukawa couplings (e.g. the top quark Yukawa coupling). Smaller Yukawa couplings are generated effectively from higher dimension non-renormalisable operators corresponding to insertions of $\theta$ and $\bar{\theta}$ fields and hence to powers of the expansion parameter in Eq.\ref{expansion}, which we have identified with the Wolfenstein parameter. The number of powers of the expansion parameter is controlled by the $U(1)_X$ charge of the particular MSSM operator. \footnote{Of course this simple picture may in reality be more complicated if several different vector mass scales are assumed, and taking into account the order one dimensionless couplings involving different $\theta$ and $\bar{\theta}$ fields coupling the MSSM fields to the heavy vector matter. By making various dynamical assumptions it is possible to generate several different expansion parameters which may be in expanded non-integer powers \cite{Ross}. It is also possible to introduce several $U(1)$ symmetries, such as a model recently proposed based on a family-independent pseudo-anomalous $U(1)_X$ symmetry together with two further anomaly-free but family-dependent $U(1)$ symmetries \cite{Ramond}. For our purposes here it is sufficient to assume a single $U(1)_X$ family symmetry with the single Wolfenstein expansion parameter in Eq.\ref{expansion} raised to integer powers.} The MSSM fields $Q_i$, $U^c_j$, $D^c_j$, $L_i$, $E^c_j$, $H_u$, $H_d$ are assigned $U(1)_X$ charges $q_i$, $u_j$, $d_j$, $l_i$, $e_j$, $h_u$, $h_d$ consistent with Eq.\ref{anomalies}. This restricts the physical values of the charges which we are permitted to assign. \footnote{This restriction may be relaxed by assuming that the heavy vector matter has $X_i$ charges chosen to cancel the anomalies, but we prefer instead to regard this as a welcome constraint on the charges. We shall, however, allow heavy MSSM singlets with arbitrary charges to cancel $U(1)_X^3$ anomalies.} We do not impose any restriction on the $Z$ right-handed MSSM singlet neutrinos $N^c_p$ which therefore have unconstrained charges $n_p$. We shall suppose that the right-handed neutrino Majorana mass matrix $M_{RR}$ arises from the VEV of another MSSM singlet $\Sigma$ with charge $\sigma$ \cite{Ross}. The anomaly restriction means that there must exist a physical basis where the Higgs charges are equal and opposite, $h_u=-h_d$ in order to cancel their contributions to the anomalies, and gives a zero charge to the $\mu H_u H_d$ term. The other operators in Eq.\ref{MSSM} will in general have non-zero charges and from Eqs.\ref{expansion}, the associated Yukawa couplings and Majorana mass terms may then be expanded in powers of the Wolfenstein parameter, \begin{eqnarray} && Y^u_{ij}\sim \lambda^{|q_i+u_j+h_u|}, \ \ Y^d_{ij}\sim \lambda^{|q_i+d_j+h_d|}, \ \ Y^e_{ij}\sim \lambda^{|l_i+e_j+h_d|}, \label{Yukexpch}\\ && Y^{\nu}_{ip} \sim \lambda^{|l_i+n_p+h_u|}, \ \ M_{RR}^{pq} \sim \lambda^{|n_p + n_q + \sigma|}<\Sigma >. \ \ \label{Yukexpneut} \end{eqnarray} In the physical basis of charges discussed so far the quarks and leptons must contribute to the anomalies in the ratios in Eq.\ref{anomalies}. A corollary of this is that the physical charges are related to traceless charges (denoted by primes) by two flavour-independent $SU(5)$ shifts $\Delta t \equiv \Delta q =\Delta u = \Delta e$ and $\Delta f \equiv \Delta l =\Delta d $ \cite{IR}: \begin{equation} q_i '=q_i+\Delta t, \ \ u_i '=u_i+ \Delta t,\ \ e_i '=e_i + \Delta t,\ \ l_i '=l_i + \Delta f, \ \ d_i '= d_i + \Delta f, \label{primes} \end{equation} It is possible to absorb the $SU(5)$ shifts into the Higgs charges by defining \begin{equation} h_u' \equiv h_u-2 \Delta t , \ \ h_d' \equiv h_d- \Delta t - \Delta f , \label{higgs'} \end{equation} so that \begin{equation} q_i+u_j+h_u=q_i'+u_j'+h_u', \ \ q_i+d_j+h_d=q_i'+d_j'+h_d', \ \ l_i+e_j+h_d=l_i'+e_j'+h_d'. \label{equivalent} \end{equation} The couplings in Eq.\ref{Yukexpch} may then be equivalently expanded in terms of primed charges. Tracelessness implies that the first family charges may be eliminated \begin{equation} q_1'= -q_2'-q_3', \ \ u_1'=-u_2'-u_3', \ \ d_1'= -d_2'-d_3', \ \ l_1'=-l_2'-l_3', \ \ e_1'= -e_2'-e_3'. \label{trace} \end{equation} Since the 33 component of the Yukawa matrices are either renormalisable or related to $\tan \beta$ dependent integers we can eliminate the primed Higgs charges using \begin{equation} h_u'=-q_3'-u_3', \ \ h_d'=n_d-q_3'-d_3'=n_e-l_3'-e_3'. \ \ \label{higgs} \end{equation} Using Eqs.\ref{equivalent}, \ref{trace}, \ref{higgs} the Yukawa matrices in \ref{Yukexpch} may then be expressed as $$ Y^u \sim \left( \begin{array}{lll} \lambda^{|\gamma_u+\delta_u|} & \lambda^{|\gamma_u+\beta_u|} & \lambda^{|\gamma_u|}\\ \lambda^{|\alpha_u+\delta_u|}& \lambda^{|\alpha_u+\beta_u|} & \lambda^{|\alpha_u|}\\ \lambda^{|\delta_u|}& \lambda^{|\beta_u|} & 1 \end{array} \right), \ \ \ \ Y^d \sim \left( \begin{array}{lll} \lambda^{|\gamma_d+\delta_d+n_d|} & \lambda^{|\gamma_d+\beta_d+n_d|} & \lambda^{|\gamma_d+n_d|}\\ \lambda^{|\alpha_d+\delta_d+n_d|}& \lambda^{|\alpha_d+\beta_d+n_d|} & \lambda^{|\alpha_d+n_d|}\\ \lambda^{|\delta_d+n_d|}& \lambda^{|\beta_d+n_d|} & \lambda^{|n_d|} \end{array} \right), $$ \begin{equation} Y^e \sim \left( \begin{array}{lll} \lambda^{|\gamma_e+\delta_e+n_e|} & \lambda^{|\gamma_e+\beta_e+n_e|} & \lambda^{|\gamma_e+n_e|}\\ \lambda^{|\alpha_e+\delta_e+n_e|}& \lambda^{|\alpha_e+\beta_e+n_e|} & \lambda^{|\alpha_e+n_e|}\\ \lambda^{|\delta_e+n_e|}& \lambda^{|\beta_e+n_e|} & \lambda^{|n_e|} \end{array} \right), \ \ \ \ \label{chargedYuks} \end{equation} where \begin{eqnarray} &&\alpha_u=\alpha_d=q_2'-q_3', \ \ \alpha_e=l_2'-l_3', \nonumber \\ &&\beta_u=u_2'-u_3',\ \ \beta_d=d_2'-d_3', \ \ \beta_e=e_2'-e_3', \nonumber\\ &&\gamma_u=\gamma_d=-q_2'-2q_3',\ \ \gamma_e=-l_2'-2l_3', \nonumber \\ &&\delta_u=-u_2'-2u_3',\ \ \delta_d=-d_2'-2d_3', \ \ \delta_e=-e_2'-2e_3'. \label{abcd} \end{eqnarray} The above analysis applies quite generally to any theory based on a single pseudo-anomalous $U(1)_X$ gauged family symmetry. However the quark and lepton charges may be constrained by imposing unification constraints on the theory. For example: \begin{itemize} \item $SU(5)$ unification implies \footnote{Note that $SU(5)$ automatically guarantees Green-Schwartz anomaly cancellation for any choice of charges.} $l_i=d_i$, $q_i=u_i=e_i$ but allows $Z$ arbitrary right-handed neutrino charges $n_p$. \item $SU(2)_R$ gauge symmetry implies that $Z=3$ with $n_i=e_i$ and $d_i=u_i$. \item Left-right symmetry is stronger than $SU(2)_R$ and implies $n_i=e_i=l_i$ and $d_i=u_i=q_i$. \item Pati-Salam $SU(4)\times SU(2)_L \times SU(2)_R$ implies $l_i=q_i$ and $u_i=d_i=e_i=n_i$. \item $SO(10)$ unification implies $l_i=d_i=q_i=u_i=e_i=n_i$ \item Trinification $SU(3)^3$ implies $u_i=d_i$, $l_i=e_i=n_i$ and unconstrained $q_i$ \item Flipped $SU(5)\times U(1)$ implies $q_i=d_i=n_i$, $u_i=l_i$ and unconstrained $e_i$ \end{itemize} As discussed in ref.\cite{Ross} these examples are difficult to reconcile with the data without either appealing to group theoretical Clebsch relations or carefully chosen dynamical assumptions. \footnote{The $SU(3)^3$ model discussed there looks the most natural.} We shall therefore not impose such gauge unification constraints here but instead consider the general case in Eqs.\ref{chargedYuks}. By comparing Eqs.\ref{chargedYuks}, to Eqs.\ref{Ramond}, \ref{Ramond'} suitable choices of the integers $\alpha_a, \beta_a, \gamma_a, \delta_a$, (where $a=u,d,e$) can readily be deduced. Note especially that $m_b/m_{\tau} \sim 1$ implies \begin{equation} n=|n_e|=|n_d|, \ \ \tan \beta \sim \lambda^{n-3}. \end{equation} It is straightforward to scan over all the possible positive and negative integers $\alpha_a, \beta_a, \gamma_a, \delta_a, n_a$ to find acceptable Yukawa matrices from Eqs.\ref{chargedYuks}. For example a special case is when $\alpha_a, \beta_a, \gamma_a, \delta_a, n_a$ are all positive definite integers \cite{Ramond}. In this case from Eqs.\ref{Ramond}, \ref{Ramond'}, \ref{chargedYuks} we find $\alpha_u=\alpha_d=2$, $\alpha_e=0$, $\beta_u=2$, $\beta_d=0$, $\beta_e=2$, $\gamma_u=\gamma_d=3$, $\gamma_e=2$, $\delta_u=5$, $\delta_d=1$, $\delta_e=2$, $n_e=n_d=n$. \footnote{ Note that $n_e=n_d=n$ imposes the non-trivial constraint that $\alpha_e+\beta_e+\gamma_e+\delta_e=\alpha_d+\beta_d+\gamma_d+\delta_d$ which is satisfied here. If for example we had taken $\frac{m_e}{m_{\tau}} \sim \lambda^5$ it would not be satisfied.} The Yukawa matrices are then fully specified in this example, up to a $\tan \beta$ dependence: \begin{equation} Y^u \sim \left( \begin{array}{lll} \lambda^8 & \lambda^5 & \lambda^3 \\ \lambda^7 & \lambda^4 & \lambda^2 \\ \lambda^5 & \lambda^2 & 1 \end{array} \right), \ \ \ \ Y^d \sim \left( \begin{array}{lll} \lambda^4 & \lambda^3 & \lambda^3 \\ \lambda^3 & \lambda^2 & \lambda^2 \\ \lambda & 1 & 1 \end{array} \right)\lambda^n, \ \ \ \ Y^e \sim \left( \begin{array}{lll} \lambda^4 & \lambda^4 & \lambda^2 \\ \lambda^2 & \lambda^2 & 1 \\ \lambda^2 & \lambda^2 & 1 \end{array} \right) \lambda^n. \end{equation} Given $\alpha_a, \beta_a, \gamma_a, \delta_a, n_a$ above and using Eqs.\ref{trace}, \ref{higgs}, \ref{abcd} we find the following traceless: \begin{eqnarray} && q'_i=\frac{1}{3}(4,1,-5), \ \ u'_i=\frac{1}{3}(8,-1,-7), \ \ d'_i=\frac{1}{3}(2,-1,-1) \nonumber \\ && l'_i=\frac{1}{3}(4,-2,-2), \ \ e'_i=\frac{1}{3}(2,2,-4), \ \ h'_u=4, \ \ h'_d=2+n \label{tracelesscharges} \end{eqnarray} The physical (unprimed) charges are by definition those which lead to the Higgs charges satisfying $h_u=-h_d$. Eq.\ref{higgs'} shows that there remains an ambiguity in the choice of Higgs charges and hence in $\Delta t , \Delta f$ which are two unknowns constrained by only one relation, namely $-3\Delta t = h'_d +h'_u +\Delta f$. We can regard $\Delta f$ as being a completely free parameter whose choice specifies all the physical (unprimed) charges uniquely. For example we may set the Higgs charges to be zero by taking \footnote{Note that in general both both $\Delta t$ and $\Delta f$ are non-zero and so the family symmetry $U(1)_X$ cannot be anomaly-free and is instead pseudo-anomalous \cite{IR}.} $\Delta t=-2$, $\Delta f = -n$ which enables the physical (unprimed) charges to be deduced from Eq.\ref{primes}. Other choices of $\Delta f$ will lead to different choices of physical charges. \section{SRHND and U(1) Family Symmetry} We now turn our attention to the neutrino sector, which is the main focus of this paper. Since the $Z$ right-handed neutrinos are not constrained by anomaly cancellation it is most convenient to work with physical (unprimed) charges as in Eq.\ref{Yukexpneut}. $Y^{\nu}$ clearly depends on the combination of lepton and Higgs charges $$ l_i+ h_u =l'_i+h'_u/3-2h'_d/3-5\Delta f /3 $$ which is not fixed by the primed charges due to the remaining freedom in $\Delta f$. In dealing with the neutrino sector it is convenient to absorb the Higgs charge $h_u$ into the definition of the lepton charges $l_i$ so that Eq.\ref{Yukexpneut} becomes \begin{equation} Y^{\nu}_{ip} \sim \lambda^{|l_i+ n_p|}, \ \ M_{RR}^{pq} \sim \lambda^{|n_p + n_q + \sigma|}<\Sigma > \ \ \label{Yukexpneut'} \end{equation} where the redefined $l_i$ are related to the traceless charges $l'_i$ by arbitrary family-independent shifts, and using Eq.\ref{tracelesscharges} may be written as: \begin{equation} l_i=(2+l_3,l_3,l_3) \label{li} \end{equation} where the numerical value of $l_3$ remains a free choice. The light Majorana matrix may then be constructed from Eq.\ref{mLL} which we repeat below \begin{equation} m_{LL}=v_2^2Y_{\nu}M_{RR}^{-1}Y_{\nu}^T \label{mLLagain} \end{equation} If we were to assume positive definite values for $l_i+n_p$ and $n_p + n_q + \sigma$ then the modulus signs could be dropped and the right-handed neutrino charges $n_p$ would cancel when $m_{LL}$ is constructed from Eqs.\ref{mLLagain} and \ref{Yukexpneut'} \cite{dropout}. The argument relies on the observation that if the modulus signs are dropped from Eq.\ref{Yukexpneut'} one can always write \begin{eqnarray} &&Y^{\nu}=diag(\lambda^{l_1},\lambda^{l_2},\lambda^{l_3})Y_D diag(\lambda^{n_1},\ldots ,\lambda^{n_Z}), \nonumber \\ &&M_{RR}=diag(\lambda^{n_1},\ldots ,\lambda^{n_Z})M_{M} diag(\lambda^{n_1},\ldots ,\lambda^{n_Z}) \label{factors} \end{eqnarray} where $Y_D$ and $M_M$ are democratic matrices. Inserting Eq.\ref{factors} into Eq.\ref{mLL} the right-handed neutrino charges are seen to cancel. Such a cancellation would imply that every right-handed neutrino would contribute equally to every entry in $m_{LL}$ regardless of the right-handed neutrino charges. From the point of view of SRHND it is therefore important that such a cancellation does not take place, and so we shall require that at least some of the combinations $l_i+n_p$ and $n_p + n_q + \sigma$ take negative values. In such a case the choice of right-handed neutrino charges will play an important role in determining $m_{LL}$, and each particular choice of $n_p$ must be analysed separately. At first sight the general case of $Z$ right-handed neutrinos with unconstrained charges $n_p$ leading to non-positive definite exponents in Eq.\ref{Yukexpneut'} seems to make the determination of $m_{LL}$ an intractable problem. However we have already argued that the atmospheric neutrino data suggests SRHND in the 23 sector and this will lead to $m_{LL}$ of the form given in Eq.\ref{mLL''}. We shall now formulate the general conditions which will lead to SRHND in the 23 sector. \subsection{One Right-handed Neutrino} Let us first consider the case $Z=1$ where there is just a single right-handed neutrino, which for later convenience we shall refer to as $N^c_3$ with charge $n_3$. In this case Eq.\ref{Yukexpneut'} becomes \begin{equation} Y^{\nu}_{i3} \sim \lambda^{|l_i+n_3|}, \ \ M_{RR}^{33} \sim \lambda^{|2n_3 + \sigma|}<\Sigma >. \ \ \label{Yukexpneut''} \end{equation} Being a $1 \times 1$ matrix $M_{RR}^{33}$ is trivially inverted and we obtain from Eqs.\ref{mLLagain}, \begin{equation} m_{LL}^{ij}\sim \lambda^{|l_i+n_3|}\lambda^{|l_j+n_3|} \frac{v_2^2}{M_{RR}^{33}} \label{mLLSRHN} \end{equation} which should be compared to Eq.\ref{matrix}, where we identify \footnote{Even though the couplings in Eq.\ref{matrix} were defined in the diagonal charged lepton basis, the identification is still valid to a consistent order of the expansion parameter.} \begin{equation} Y^{\nu}_{i3} \sim \lambda^{|l_i+n_3|} \sim (\lambda_{\nu_e}, \lambda_{\nu_{\mu}}, \lambda_{\nu_{\tau}}) \end{equation} Then Eq.\ref{mLL''} requires that \begin{equation} |l_2+n_3|=|l_3+n_3|, \ \ \ \ |l_1+n_3|-|l_3+n_3|=2 \label{modulus} \end{equation} If both $l_2+n_3$ and $l_3+n_3$ have the same sign (SS) then $l_2=l_3$, whereas if they have opposite signs (OS) then $l_2+l_3=-2n_3$. Similarly if both $l_1+n_3$ and $l_3+n_3$ have the SS then $l_1-l_3=2$, whereas if they have OS then $l_1+l_3=2-2n_3$. Interestingly the SS cases $l_2=l_3$, $l_1-l_3=2$ have already arisen in the example in Eq.\ref{tracelesscharges}, which corresponds to $l_i$ charges in Eq.\ref{li}. This is no surprise since it originates from the charged lepton Yukawa matrix in Eq.\ref{Ramond'} which follows from the assumption $V_{MNS}\sim V_{\nu L} \sim V_{eL}$ and the Super-Kamiokande data and the MSW solution. To summarise, from Eqs.\ref{Yukexpneut''}, \ref{mLLSRHN} and imposing Eq.\ref{modulus} the single right-handed neutrino included so far leads to \begin{equation} m_{LL} \sim \left( \begin{array}{lll} \lambda^4 & \lambda^{2} & \lambda^2 \\ \lambda^2 & 1 & 1 \\ \lambda^2 & 1 & 1 \end{array} \right) m_{\nu_3} \label{mLL'''} \end{equation} where the atmospheric neutrino mass is given \begin{equation} m_{\nu_3}\sim \lambda^{2|l_3+n_3|-|2n_3 +\sigma |} \frac{v_2^2}{<\Sigma >} \end{equation} With only a single right-handed neutrino $m_{LL}$ in Eq.\ref{mLL'''} has two zero eigenvalues, and a vanishing determinant of the 23 submatrix, as in Eq.\ref{matrix}. In order to implement the small angle MSW solution we need to include the effect of subdominant right-handed neutrinos which break the massless degeneracy. SRHND requires that the elements in the 23 sector of Eq.\ref{mLL'''} must receive corrections of order $\lambda^2$ from the subdominant neutrinos so that the determinant of the 23 submatrix only {\em approximately} vanishes to this order leading to a small eigenvalue of order $\lambda^2$ and the desired mass hierarchy in Eq.\ref{23hierarchy}. \subsection{Two Right-handed Neutrinos} We now include a second right-handed neutrino $N_2^c$ with charge $n_2$, in addition to $N_3^c$ with charge $n_3$. With two right-handed neutrinos, $Z=2$, the heavy Majorana mass matrix from Eq.\ref{Yukexpneut'} is \begin{equation} M_{RR} \sim \left( \begin{array}{ll} \lambda^{|2n_2+\sigma|} & \lambda^{|n_2+n_3+\sigma|} \\ \lambda^{|n_2+n_3+\sigma|} & \lambda^{|2n_3+\sigma|} \end{array} \right) <\Sigma > \label{mRR2} \end{equation} For SRHND we clearly require $n_2 \neq n_3$ to avoid the two right-handed neutrinos contributing democratically. More generally for SRHND we need to avoid large right-handed neutrino mixing angles. If we assume without loss of generality that $\lambda^{|2n_2+\sigma|} > \lambda^{|2n_3+\sigma|}$, so that $N_2^c$ is heavier than $N_3^c$, then this implies \begin{equation} |2n_2+\sigma| < |2n_3+\sigma| \label{c1} \end{equation} Then the small mixing angle requirement is \begin{equation} |2n_2+\sigma| < |n_2+n_3+\sigma| \label{c2} \end{equation} The lightest eigenvalue is of order the diagonal element provided \begin{equation} |2n_2+\sigma| \leq 2|n_2+n_3+\sigma|- |2n_3+\sigma| \label{c3} \end{equation} Assuming all these conditions are met then $M_{RR}$ will be diagonalised by small angle rotations and have hierarchical eigenvalues set by the diagonal elements. As a first approximation we may drop the off-diagonal elements and write \begin{equation} M_{RR}\approx diag(M_{R2}, M_{R3}) \label{diag} \end{equation} where \begin{equation} M_{R2} \sim \lambda^{|2n_2+\sigma|}<\Sigma >, \ \ M_{R3} \sim \lambda^{|2n_3+\sigma|}<\Sigma > \label{MR23} \end{equation} Then the light Majorana matrix is given by adding the separate contribution from each of the two right-handed neutrinos \begin{equation} m_{LL}^{ij}=v_2^2\left( \frac{{{Y}}_{\nu}^{i2} {{Y}}_{\nu}^{j2 } }{M_{R2}}+ \frac{{{Y}}_{\nu}^{i3} {{Y}}_{\nu}^{j3} }{M_{R3}} \right) \end{equation} It is clear that the dominant contribution to a particular element of $m_{LL}$ will come from the right-handed neutrino which is at the same time the lightest, and couples the most strongly to left-handed neutrinos. Without loss of generality we have taken $N_3^c$ to be the lighter right-handed neutrino and to give the dominant contribution to the 23 block of $m_{LL}$ in Eq.\ref{mLLSRHN}. We therefore write the subdominant contribution coming from the second right-handed neutrino $N_2^c$ as \begin{equation} \delta m_{LL}^{ij} =\lambda^{|l_i+n_2|}\lambda^{|l_j+n_2|} \frac{v_2^2}{M_{R2}} \end{equation} As discussed below Eq.\ref{mLL'''} we require: \begin{equation} \frac{\delta m_{LL}^{33}}{m_{LL}^{33}} = \frac{\lambda^{2|l_3+n_2|} }{\lambda^{2|l_3+n_3|} } \frac{M_{R3}}{M_{R2}} \sim \lambda^2 \label{cond} \end{equation} From Eqs.\ref{MR23}, \begin{equation} \frac{M_{R3}}{M_{R2}} \sim \lambda^{|2n_3+\sigma|-|2n_2+\sigma|} \end{equation} so Eq.\ref{cond} implies the condition \begin{equation} 2|l_3+n_2|-2|l_3+n_3|+|2n_3+\sigma|-|2n_2+\sigma|=2 \label{condition} \end{equation} We already observed that the required MSW perturbation is \begin{equation} \delta m_{LL}^{33} \sim \frac{v_2^2}{M_U} \end{equation} so we deduce \begin{equation} \frac{M_{R2}}{M_U}\sim \lambda^{2|l_3+n_2|}, \ \ \frac{<\Sigma >}{M_U}\sim \lambda^{2|l_3+n_2|- |2n_2+\sigma|} \end{equation} There is the further requirement that the powers of $\lambda$ occuring in $M_{RR}$ and $Y_{\nu}$ be either integer or half-integer. \footnote{In the case of half-integer powers this implies that the $\theta$, $\bar{\theta}$ fields which break the $U(1)_X$ symmetry must have charges $\pm 1/2$ and the expansion parameter in Eq.\ref{expansion} must be redefined so that $\frac{<\theta >}{M_V}=\frac{<\bar{\theta} >}{M_V}= \lambda^{1/2}$, as in ref.\cite{Ross}.} By scanning over half-integer and integer values of $l_3,n_2,n_3,\sigma$ we find that there are no solutions which satisfy all the above constraints for integer powers of $\lambda$ in $M_{RR}$ and $Y_{\nu}$.\footnote{I am grateful to Y. Nir (private communication) for pointing this out.} However there are a large number of solutions involving half-integer powers of $\lambda$ in $M_{RR}$ and $Y_{\nu}$ (of course $m_{LL}$ in Eq.\ref{mLL''} always involves integer powers of $\lambda$.) The condition in Eq.\ref{condition} may be achieved in various ways with $N_3^c$ being lighter than $N_2^c$ by a factor, $M_{R3}/M_{R2} \sim \lambda^{|2n_3+\sigma|-|2n_2+\sigma|}\sim \lambda^a$, and the ratio of the Dirac couplings of $N_2^c$, $N_3^c$ to $L_3$ given by $\lambda^{{2|l_3+n_2|}- 2|l_3+n_3|} \sim \lambda^{2-a}$, where $a>0$ is a positive integer. For example $l_3=-1/2,n_2=0,n_3=1,\sigma =0$ satisfies all the conditions with $a=2$ and $Y_{\nu}$ involving half-integer exponents. Further examples are listed in Table 1. \begin{table}[tbp] \hfil \begin{tabular}{ccccccc} \hline $l_3$ & $n_2$& $n_3$ & $\sigma$ & $a$ \\ \hline -1 & -1/2 & 0 & 1 & 1 \\ \hline -1 & 0 & 1/2 & 0 & 1 \\ \hline -1 & 0 & 1/2 & 1/2 & 1 \\ \hline -1 & 0 & 1/2 & 1 & 1 \\ \hline -1 & 1/2 & 1 & -1 & 1 \\ \hline -1 & 1/2 & 1 & -1/2 & 1 \\ \hline -1 & 1/2 & 1 & 0 & 1 \\ \hline -1 & 1/2 & 1 & 1/2 & 1 \\ \hline -1 & 1/2 & 1 & 1 & 1 \\ \hline -1/2 & -1/2 & 0 & 1 & 1 \\ \hline -1/2 & 0 & 1/2 & 0 & 1 \\ \hline -1/2 & 0 & 1/2 & 1/2 & 1 \\ \hline -1/2 & 0 & 1/2 & 1 & 1 \\ \hline 0 & -1/2 & 0 & 1 & 1 \\ \hline 0 & 1/2 & 0 & -1 & 1 \\ \hline 1/2 & 0 & -1/2 & -1 & 1 \\ \hline 1/2 & 0 & -1/2 & -1/2 & 1 \\ \hline 1/2 & 0 & -1/2 & 0 & 1 \\ \hline 1/2 & 1/2 & 0 & -1 & 1 \\ \hline 1 & -1/2 & -1 & -1 & 1 \\ \hline 1 & -1/2 & -1 & -1/2 & 1 \\ \hline 1 & -1/2 & -1 & 0 & 1 \\ \hline 1 & -1/2 & -1 & 1/2 & 1 \\ \hline 1 & -1/2 & -1 & 1 & 1 \\ \hline 1 & 0 & -1/2 & -1 & 1 \\ \hline 1 & 0 & -1/2 & -1/2 & 1 \\ \hline 1 & 0 & -1/2 & 0 & 1 \\ \hline 1 & 1/2 & 0 & -1 & 1 \\ \hline -1/2 & 0 & 1 & 0 & 2 \\ \hline -1/2 & 0 & 1 & 1/2 & 2 \\ \hline -1/2 & 0 & 1 & 1 & 2 \\ \hline 0 & -1/2 & 1/2 & 1 & 2 \\ \hline 0 & 1/2 & -1/2 & -1 & 2 \\ \hline 1/2 & 0 & -1 & -1 & 2 \\ \hline 1/2 & 0 & -1 & -1/2 & 2 \\ \hline 1/2 & 0 & -1 & 0 & 2 \\ \hline 0 & -1/2 & 1 & 1 & 3 \\ \hline 0 & 1/2 & -1 & -1 & 3 \\ \hline \end{tabular} \hfil \caption{\footnotesize Simple $Z=2$ examples which satisfy all the conditions of SRHND given in the text.} \end{table} \subsection{Three Right-handed Neutrinos} We now wish to extend the discussion to include three right-handed neutrinos $Z=3$, by introducing a third right-handed neutrino $N_1^c$ with charge $n_1$ in addition to the two already introduced above. Again we shall suppose that $N_3^c$ gives the dominant contribution to the 23 sector masses. As for the $Z=2$ case we require $n_3 \neq n_2,n_1$, and we need to ensure that $N_3^c$ does not have large mixing angles in $M_{RR}$ in order to isolate it from the other right-handed neutrinos. This can be ensured by a sequence of conditions similar to Eqs.\ref{c1}, \ref{c2}, \ref{c3}. Then, after small angle rotations, $M_{RR}$ can be written in block diagonal form. \begin{equation} M_{RR} \sim \left( \begin{array}{lll} \lambda^{|2n_1+\sigma|} & \lambda^{|n_1+n_2+\sigma|} & 0 \\ \lambda^{|n_2+n_1+\sigma|} & \lambda^{|2n_2+\sigma|} & 0 \\ 0 & 0 & \lambda^{|2n_3+\sigma|} \\ \end{array} \right) <\Sigma > \label{mRR3} \end{equation} which is the analogue of Eq.\ref{diag}. The new feature of the $Z=3$ case compared to the $Z=2$ case is that there are now several possibilities for the structure of the upper $2 \times 2$ block in Eq.\ref{mRR3} which are all consistent with SRHND, which are listed below. ``Diagonal dominated'' corresponding to $|n_1+n_2+\sigma|>min(|2n_1+\sigma|,|2n_2+\sigma|)$: \begin{equation} M_{RR}^{upper} \sim \left( \begin{array}{ll} \lambda^{|2n_1+\sigma|} & 0 \\ 0 & \lambda^{|2n_2+\sigma|} \end{array} \right) <\Sigma > \label{mRR2a} \end{equation} ``Off-diagonal dominated'' corresponding to $|n_1+n_2+\sigma|<|2n_1+\sigma|,|2n_2+\sigma|$: \begin{equation} M_{RR}^{upper} \sim \left( \begin{array}{ll} 0 & \lambda^{|n_1+n_2+\sigma|} \\ \lambda^{|n_2+n_1+\sigma|} & 0 \end{array} \right) <\Sigma > \label{mRR2b} \end{equation} ``Democratic'' corresponding to $|n_1+n_2+\sigma| = |2n_1+\sigma| = |2n_2+\sigma|$: \begin{equation} M_{RR}^{upper} \sim \left( \begin{array}{ll} \lambda^{|2n_1+\sigma|} & \lambda^{|n_1+n_2+\sigma|} \\ \lambda^{|n_2+n_1+\sigma|} & \lambda^{|2n_2+\sigma|} \end{array} \right) <\Sigma > \label{mRR2c} \end{equation} In the ``diagonal dominated'' case after small angle rotations the light effective Majorana mass matrix in Eq.\ref{mLLagain} may be calculated in the diagonal right-handed neutrino basis \begin{equation} m_{LL}=v_2^2Y_{\nu}M_{RR}^{-1}Y_{\nu}^T =v_2^2Y_{\nu}\Omega_{RR}^{\dag}(M_{RR}^{diag})^{-1} \Omega_{RR}Y_{\nu}^T \label{mLL''''} \end{equation} The advantage of working in a diagonal right-handed neutrino mass basis is that $(M_{RR}^{diag})^{-1} ={\rm diag(M_{R1}^{-1},M_{R2}^{-1},M_{R3}^{-1})}$ so if we define ${\tilde{Y}}_{\nu} \equiv Y_{\nu}\Omega_{RR}^{\dag}$ as the neutrino Yukawa matrix in the diagonal right-handed neutrino basis, then the effective light mass matrix elements are given from Eq.\ref{mLL''''} by \begin{equation} m_{LL}^{ij}=\sum_{p=1}^3 v_2^2\frac{{\tilde{Y}}_{\nu}^{ip} {\tilde{Y}}_{\nu}^{jp } }{M_{Rp}} \label{mLL'} \end{equation} In this case $\Omega_{RR}$ involves small angle rotations and so $\tilde{Y}_{\nu} \sim Y_{\nu}$, and the contributions to $m_{LL}$ from the neutrinos $N_1^c,N_2^c$ are: \begin{equation} \delta m_{LL}^{ij}= v_2^2\left( \frac{{{Y}}_{\nu}^{i1} {{Y}}_{\nu}^{j1 } }{M_{R1}}+ \frac{{{Y}}_{\nu}^{i2} {{Y}}_{\nu}^{j2} }{M_{R2}} \right) \end{equation} where \begin{equation} M_{R1} \sim \lambda^{|2n_1+\sigma|}<\Sigma >, \ \ M_{R2} \sim \lambda^{|2n_2+\sigma|}<\Sigma > \label{MR1MR2} \end{equation} and from Eq.\ref{Yukexpneut'} $Y_{\nu}^{ip}=\lambda^{|l_i+n_p|}$. Similar to Eq.\ref{cond} in this case we require \begin{equation} \frac{\delta m_{LL}^{33}}{m_{LL}^{33}} \sim \frac{\lambda^{2|l_3+n_1|} }{\lambda^{2|l_3+n_3|} } \frac{M_{R3}}{M_{R1}} + \frac{\lambda^{2|l_3+n_2|} }{\lambda^{2|l_3+n_3|} } \frac{M_{R3}}{M_{R2}} \sim \lambda^2 \label{conda} \end{equation} Thus the conditions for the ``diagonal dominated case'' are: \begin{eqnarray} && 2|l_3+n_1|-2|l_3+n_3| + |2n_3+\sigma| -|2n_1+\sigma|\geq 2, \nonumber \\ && 2|l_3+n_2|-2|l_3+n_3|+ |2n_3+\sigma|-|2n_2+\sigma| \geq 2 \label{a} \end{eqnarray} where at least one of the inequalities must be saturated. In the ``off-diagonal dominated'' case $M_{RR}$ can again be simply inverted leading to \begin{equation} \delta m_{LL}^{ij}= v_2^2\left( \frac{{{Y}}_{\nu}^{i1} {{Y}}_{\nu}^{j2 } }{M_{R12}}+ \frac{{{Y}}_{\nu}^{i2} {{Y}}_{\nu}^{j1} }{M_{R12}} \right) \end{equation} where \begin{equation} M_{R12} \sim \lambda^{|n_1+n_2+\sigma|}<\Sigma >. \label{MR12} \end{equation} Again similar to Eq.\ref{cond} we require \begin{equation} \frac{\delta m_{LL}^{33}}{m_{LL}^{33}} \sim \frac{\lambda^{|l_3+n_1|+|l_3+n_2|} }{\lambda^{2|l_3+n_3|} } \frac{M_{R3}}{M_{R12}} \sim \lambda^2 \label{condb} \end{equation} Thus the condition for the ``off-diagonal dominated case'' is: \begin{equation} |l_3+n_1|+|l_3+n_2|-2|l_3+n_3| + |2n_3+\sigma| -|n_1+n_2+ \sigma|= 2. \label{b} \end{equation} In the ``democratic'' case $M_{RR}$ can be readily inverted leading to a result of order \begin{equation} \delta m_{LL}^{ij}\sim v_2^2\left( \frac{{{Y}}_{\nu}^{i1} {{Y}}_{\nu}^{j1 } }{M}\right) \end{equation} where the right-handed neutrino masses in the upper block, $M$, are all equal by the democratic assumption and we have specialised to $n_1=n_2$ which implies from Eq.\ref{Yukexpneut'} that ${Y}_{\nu}^{i1}\sim {Y}_{\nu}^{i2}$. Once again similar to Eq.\ref{cond} we require \begin{equation} \frac{\delta m_{LL}^{33}}{m_{LL}^{33}} \sim \frac{\lambda^{2|l_3+n_1|} }{\lambda^{2|l_3+n_3|} } \frac{M_{R3}}{M} \sim \lambda^2 \label{condc} \end{equation} Thus the condition for the ``democratic case'' is: \begin{equation} 2|l_3+n_1| -2|l_3+n_3| + |2n_3+\sigma| -|2n_1+ \sigma|= 2. \label{c} \end{equation} In practice examples of all three kinds can easily be constructed along the same lines as the explicit $Z=2$ case. The ``democratic'' case with $n_1=n_2$ is isomorphic to the $Z=2$ case. The $Z=2$ results trivially generalise in this case to $n_p=(n_2,n_2,n_3)$ where some examples of charges were listed in Table 1. For example $l_3=-1/2, n_p=(0,0,1), \sigma =0$ satisfies all the ``democratic'' conditions with $a=2$ and $Y_{\nu}$ involving half-integer exponents. Clearly in the ``democratic'' case the $Z=2$ results can immediately be generalised to any number of right-handed neutrinos $Z$ with $n_p=(n_2, \ldots, n_2,n_3)$, where $(n_2, n_3)$ are the $Z=2$ charges. The ``diagonal dominated'' case also follows a similar pattern to the $Z=2$ case with the lighter of $N_1^c$, $N_2^c$ playing the role of the subdominant right-handed neutrino in the $Z=2$ case. It is straightforward to scan over all the half-integer and integer charges which satisfy the ``diagonal dominated'' conditions and generate a list of charges for this case, analagous to Table 1. A single example will suffice: $l_3=-3/2, n_p=(0,1,2), \sigma =0$ satisfies all the ``diagonal dominated'' conditions and Eq.\ref{a} is saturated by $N_2^c$ which plays the role of the subdominant right-handed neutrino of the $Z=2$ case, with $N_1^c$ being both heavier and having more suppressed Dirac couplings. Again the ``diagonal dominated'' case can immediately be generalised to any number $Z$ of right-handed neutrinos $n_p=(n_q, n_2,n_3)$, where $(n_2, n_3)$ are the $Z=2$ charges with $N_2^c$ playing the role of the subdominant right-handed neutrino and $N_q^c$ giving subsubdominant contributions to the 23 block of $m_{LL}$. Examples of the ``off-diagonal dominated'' kind have already been proposed in the literature, although they were not interpreted as being due to SRHND \cite{Altarelli}. To show that the models in ref.\cite{Altarelli} are examples of SRHND of the ``off-diagonal dominated'' kind it suffices to consider a specific example: \begin{equation} l_i=(2,0,0),\ \ n_p=(1,-1,0), \ \ \sigma = 0 \label{Altcharges} \end{equation} It is immediately clear that the charges in Eq.\ref{Altcharges} satisfy the conditions for SRHND in general Eq.\ref{modulus} and in particular the ``off-diagonal dominated'' conditions $|n_1+n_2+\sigma|<|2n_1+\sigma|,|2n_2+\sigma|$ and Eq.\ref{b}. This immediately substantiates our claim that these models correspond to SRHND of the ``off-diagonal dominated'' kind. Note that $Y^{\nu}$ involves integer exponents. In view of the interest in this example in the literature we develop it in a little more detail below. The charges in Eq.\ref{Altcharges} lead to the following neutrino Yukawa and heavy Majorana matrices \begin{equation} Y^{\nu} \sim \left( \begin{array}{lll} \lambda^3 & \lambda & \lambda^2 \\ \lambda & \lambda & 1 \\ \lambda & \lambda & 1 \end{array} \right), \ \ \ \ M_{RR} \sim \left( \begin{array}{lll} \lambda^2 & 1 & \lambda \\ 1 & \lambda^2 & \lambda \\ \lambda & \lambda & 1 \end{array} \right) <\Sigma > \end{equation} Due to the assumed charges, the heavy Majorana matrix is dominated by three equal mass terms $<\Sigma >N^c_1N^c_2$, $<\Sigma >N^c_2N^c_1$ and $<\Sigma >N^c_3N^c_3$, leading to three roughly degenerate right-handed neutrinos. However of the three right-handed neutrinos it is $N^c_3$ which couples dominantly to the left-handed neutrinos of the second and third family, due to the assumed choice of $X$ charges, and hence dominates the 23 sector of $m_{LL}$. To see this we evaluate $m_{LL}$ in the basis in which \begin{equation} M_{RR} \sim \left( \begin{array}{lll} 0 & 1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 1 \end{array} \right) <\Sigma >, \ \ \ \ M_{RR}^{-1} \sim \left( \begin{array}{lll} 0 & 1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 1 \end{array} \right) <\Sigma >^{-1} \label{basis} \end{equation} In this basis we define ${\tilde{Y}}_{\nu} \equiv Y_{\nu}\Omega_{RR}^{\dag}$ where \begin{equation} \Omega_{RR} \sim \left( \begin{array}{lll} 1 & \lambda^2 & \lambda \\ \lambda^2 & 1 & \lambda \\ \lambda & \lambda & 1 \end{array} \right) \end{equation} Evaluating $m_{LL}$ in this basis we find from Eqs.\ref{mLL} and \ref{basis} \begin{equation} m_{LL}^{ij}= \frac{v_2^2}{<\Sigma >} (\tilde{Y}_{\nu}^{i1} {\tilde{Y}}_{\nu}^{j2 } + \tilde{Y}_{\nu}^{i2} {\tilde{Y}}_{\nu}^{j1 } +\tilde{Y}_{\nu}^{i3} {\tilde{Y}}_{\nu}^{j3 }) \label{mLL '''} \end{equation} corresponding to the contributions from the inverse mass terms $<\Sigma >N^c_1N^c_2$, $<\Sigma >N^c_2N^c_1$ and $<\Sigma >N^c_3N^c_3$, respectively. Since ${\tilde{Y}}_{\nu} \sim Y_{\nu}$ with the order one contributions to ${\tilde{Y}}_{\nu}$ coming exclusively from $N^c_3$, it is clear (by explicit evaluation of Eq.\ref{mLL '''}) that $N^c_3$ dominates the contributions to the 23 block of $m_{LL}$, with corrections of order $\lambda^2$ coming from the other contributions. The remaining parts of $m_{LL}$ receive contributions at the same order as the $N^c_3$ contributions coming from $N^c_1 , N^c_2$. Thus the resulting light effective neutrino matrix is as in Eq.\ref{mLL''}, with SRHND in the 23 sector due to $N^c_3$ dominance with $O(\lambda^2)$ corrections from other right-handed neutrinos. Finally we note that for $Z>3$ the above three categories ``diagonal dominated'', ``off-diagonal dominated'' and ``democratic'' may be combined in all possible ways. \section{Conclusion} We have suggested a natural explanation of both neutrino mass hierarchies {\it and} large neutrino mixing angles, as required by the atmospheric neutrino data, in terms of a single right-handed neutrino giving the dominant contribution to the 23 block of the light effective neutrino matrix. We illustrated this mechanism in the framework of models with a single pseudo-anomalous $U(1)_X$ family symmetry, expanding all masses and mixing angles in terms of the Wolfenstein parameter $\lambda$. Sub-dominant contributions to the 23 sector from other right-handed neutrinos, suppressed by a factor of $\lambda^2$, are required to give small mass splittings appropriate to the small angle MSW solution to the solar neutrino problem. We gave general conditions for achieving this in the framework of $U(1)_X$ family symmetry models containing arbitrary numbers of right-handed neutrinos $Z$. We classified the $Z=3$ cases into three categories: ``diagonal dominated'', ``off-diagonal dominated'' and ``democratic'', and discussed examples of each kind. Although the approach in \cite{Altarelli} is based on the formal condition that the subdeterminant vanishes to order $\lambda^2$, we have shown that explicit examples of this kind of model may be classified within our framework as SRHND of the ``off-diagonal dominated'' kind. Although we discussed a particular family symmetry it is clear that the idea of SRHND is more general and has recently been used in a model with $SU(2)$ family symmetry \cite{Barbieri}.
\section{Introduction} In recent proposals to perform quantum computation in nuclear spin systems using nuclear magnetic resonance (NMR) techniques~\cite{Gershenfeld97,Cory97x,Cory97b,Chuang97e}, coupled logic operations are performed using spin-spin couplings that occur naturally in molecular systems. While this is straightforward for small systems with a few spins, generalization to complex molecular structures has been challenging. The complication is caused by the many spin-spin couplings which occur along with the desired one. This fundamental task to turn off spurious evolution is so difficult that, coercing a complex system to {\em do nothing}~\cite{Jozsa98} -- ceasing all evolution -- can be just as difficult as making it do something computationally useful. The task of turning off all couplings is known in the art of NMR as {\em decoupling}; doing this for all but a select subset of couplings is known as {\em selective recoupling}. A common method to perform these tasks is to interrupt the free evolution by carefully chosen pulses. These pulses are single spin-1/2 (qubit) operations that transform the hamiltonian in the time between pulses in such a manner that unwanted evolutions in consecutive time intervals cancel out each other. Pulse sequences which perform selective recoupling are generally difficult to find for a large system. Each pulse simultaneously affects many coupling terms in the hamiltonian. To turn off all but one of the coupling terms, these pulses have to satisfy many simultaneous requirements. Ingenious sequences have been found for usual NMR applications~\cite{Ernst94,Slichter,Mansfield} but they do not address the problems relevant to quantum computation. In usual NMR applications, the structure of the spin systems is not known a-priori. Therefore, pulse sequences are designed to address all the spins together rather than individual spins. Quantum computation brings new requirements, and initial efforts~\cite{Linden98} have been made to develop pulse sequences to satisfy these needs; however, to-date, schemes have necessitated resources (such as total number of pulses applied) exponential in the number of spins being controlled. In this paper, we present an {\em efficient} scheme to perform selective recoupling. In contrast to the situation with traditional NMR, this scheme addresses the problem relevant to NMR quantum computation, in which the molecular structure is assumed to be well-known, and spins are individually addressible. The method is related to a class of well-known matrices called {\em Hadamard matrices}. We derive from any $n \times n$ Hadamard matrix a pulse scheme that decouples $n$ spins using $n$ time intervals and ${\cal O}(n^2)$ pulses. This decoupling scheme can easily be modified (i) to remove Zeeman evolution and (ii) to implement selective recoupling. This completes the construction for $n$ spins whenever $n \times n$ Hadamard matrices exist. When $n \times n$ Hadamard matrices do not exist, a scheme for $n$ spins can still be constructed using larger existing Hadamard matrices. In doing so, an extra amount of effort is required, but this can be bounded using existence properties of Hadamard matrices. Altogether, the scheme requires $cn$ time intervals and less than $cn^2$ pulses, where $c \approx 1$ with upper bound $c \leq 2$. Our method applies whenever the spins couple pairwise and have very different Zeeman frequencies, such as in heteronuclear spin systems. The paper is structured as follows. In Section~\ref{sec:statement}, we review relevant concepts in NMR quantum computing and re-state the problem precisely. In Section~\ref{sec:scheme}, we first motivate the construction of the decoupling scheme with examples, and then describe the general construction related to Hadamard matrices. Important properties of Hadamard matrices are summarized. Modifications of the decoupling scheme to perform selective recoupling are described. We conclude with some general remarks and discussions of various properties and limitations of the scheme. \section{NMR quantum computing and the statement of the problem} \label{sec:statement} In this section, we describe the NMR system and describe how a universal set of (non-fault tolerant) operations~\cite{Div95,Barenco95,Deutsch95}, namely, the single qubit operations and the controlled-NOT gate, can be realized using basic NMR primitives. We shall consider a physical system which consists of a solution of identical molecules. Each molecule has $n$ non-magnetically equivalent nuclear spins which serve as qubits. A static magnetic field is applied externally along the $+\hat{z}$ direction. This magnetic field splits the energy levels of the spin states aligned with and against it. This is described in the hamiltonian by the Zeeman terms, which, in the energy eigenbasis, are given by \begin{equation} {\cal H}_{\rm Z} = -\frac{1}{2} \sum_i \hbar \omega_i \sigma_z^{(i)} \,, \label{eq:ham:zeeman} \end{equation} where $i$ is the spin index, $\omega_i/2\pi$ is the {\em Zeeman frequency} for the $i$-th spin, and $\sigma_z^{(i)}$ is the Pauli matrix operating on the $i$-th spin. The convention $\hbar = 1$ is used for the rest of the paper. The spins have very different Zeeman frequencies, a situation loosely termed as ``heteronuclear'' in this paper. Nuclear spins can interact via the dipolar coupling~\cite{Slichter,Abragam}. This is given by the hamiltonian \begin{equation} {\cal H}_{\rm d} = \sum_{i<j} g_{ij}^{\rm d} \left[\rule{0pt}{2.4ex} \vec{\sigma}^{(i)} \cdot \vec{\sigma}^{(j)} - 3 (\hat{r}_{ij} \cdot \vec{\sigma}^{(i)}) \otimes (\hat{r}_{ij} \cdot \vec{\sigma}^{(j)}) \right] \label{eq:dipolarfull} \,, \end{equation} where $\hat{r}_{ij}$ denotes the unit displacement vector from the $i$-th to the $j$-th spin, and $g_{ij}^{\rm d}$ denotes the coupling constant between them. Spin-spin coupling can also be mediated by coupling to electrons. This indirect coupling has a tensor part and a scalar part. The tensor part is usually of the same form as ${\cal H}_{\rm d}$. The scalar part is given by the hamiltonian \begin{equation} {\cal H}_{\rm s} = \sum_{i<j} g_{ij}^{\rm s} \vec{\sigma}^{(i)} \cdot \vec{\sigma}^{(j)} \label{eq:scalar} \,. \end{equation} If the molecules tumble fast and isotropically, dipolar coupling and indirect tensor coupling will be averaged away; otherwise, the physics can be more complicated. However, in the presence of a strong external magnetic field, only the {\em secular} part (terms that commute with ${\cal H}_{\rm Z}$) is important~\cite{Slichter,Abragam}. For a heteronuclear system, the resulting coupling becomes \begin{equation} {\cal H}_{\rm c} = \sum_{i<j} g_{ij}^{\rm c} \sigma_z^{(i)} \otimes \sigma_z^{(j)} \label{eq:dipolar} \,, \end{equation} independent of the original form of coupling. Single qubit operations are performed by applying {\em pulsed} radio frequency (RF) magnetic fields along some directions $\hat{\eta}$ perpendicular to the static field. To address the $i$-th spin, the frequency of the RF field is chosen to approximate $\omega_i/2\pi$. When the $\omega_i$'s are very different, very short pulses can be used, so that during the pulses, all other evolutions are negligible except for the rotation operator $e^{-i \frac{\theta}{2} \vec{\sigma}^{(i)} \cdot \hat{\eta}}$ where $\theta$ is proportional to the pulse duration and the power. The Lie group of all single qubit operations can be generated by rotations about $\hat{x}$ and $\hat{y}$. Our scheme uses only rotations of $\theta = \pi$ along $\hat{x}$, which implement $\sigma_x$ (up to an irrelevant overall phase) on the spins being addressed. We denote this operation by $X$, superscripted by the spin index whenever appropriate. Coupled operations such as controlled-phase-shift or controlled-NOT acting on the $i$-th and the $j$-th spins can be performed given the primitive, \begin{equation} Z\!\!Z_{ij} = e^{-i \frac{\pi}{4} \sigma_z^{(i)} \otimes \sigma_z^{(j)}} \label{eq:zzij} \,. \end{equation} For instance, a controlled-NOT from the $i$-th to the $j$-th spin can be implemented by \begin{equation} e^{-i \frac{\pi}{4} \sigma_y^{(i)}} e^{i \frac{\pi}{4} \sigma_x^{(i)}} e^{i \frac{\pi}{4} \sigma_y^{(i)}} e^{-i \frac{\pi}{4} \sigma_x^{(j)}} e^{i \frac{\pi}{4} \sigma_y^{(j)}} Z\!\!Z_{ij} e^{-i \frac{\pi}{4} \sigma_y^{(j)}} \,. \end{equation} The ultimate goal is to be able to efficiently realize arbitrary quantum operations on an $n$ spin system with arbitrary couplings. In this paper, we consider a more limited objective, which can now be stated precisely, using the definitions of Eq.(\ref{eq:ham:zeeman}), Eq.(\ref{eq:dipolar}), and Eq.(\ref{eq:zzij}): \begin{quote} Given a heteronuclear system of $n$ spins with free evolution $e^{-i ({\cal H}_{\rm Z} + {\cal H}_{\rm c}) t}$, controlled using typical RF pulses, how can $Z\!\!Z_{ij}$ be implemented efficiently? \end{quote} \section{Construction of the scheme} \label{sec:scheme} We will first construct a decoupling scheme to remove the entire coupling term ${\cal H}_{\rm c}$ in the total evolution. The scheme is derived from Hadamard matrices, which will be reviewed. Methods to remove ${\cal H}_{\rm Z}$ and to implement selective recoupling will be described afterwards. \subsection{Construction of the decoupling scheme} \label{sec:examples} To construct the decoupling scheme, we only consider ${\cal H}_{\rm c}$ in the evolution. Effects of ${\cal H}_{\rm Z}$ can be included later since all matrix exponents commute. To motivate the general construction, we analyze the simplest example of decoupling two spins. The evolution operator for an arbitrary duration $t$ is given by $\tau = e^{-i g_{12}^{\rm c} t \sigma_z^{(1)} \otimes \sigma_z^{(2)}}$. The sequence of events, $\tau X^{(2)} \tau X^{(2)}$, known as {\em refocusing} in NMR, will first couple and then decouple the spins. This can be described mathematically as: \begin{eqnarray} & & \tau (X^{(2)} \tau X^{(2)}) \label{eq:refocus} \\ & = & e^{-i \theta \sigma_z^{(1)} \otimes \sigma_z^{(2)}} (\sigma_x^{(2)} e^{-i \theta \sigma_z^{(1)} \otimes \sigma_z^{(2)}} \sigma_x^{(2)}) \\ & = & e^{-i \theta \sigma_z^{(1)} \otimes \sigma_z^{(2)}} e^{-i \theta \sigma_z^{(1)} \otimes \sigma_x^{(2)} \sigma_z^{(2)} \sigma_x^{(2)}} \label{eq:third} \\ & = & e^{-i \theta \sigma_z^{(1)} \otimes \sigma_z^{(2)}} e^{-i \theta \sigma_z^{(1)} \otimes (-\sigma_z^{(2)})} \label{eq:fourth} \\ & = & e^{-i \theta \sigma_z^{(1)} \otimes \sigma_z^{(2)}} e^{i \theta \sigma_z^{(1)} \otimes \sigma_z^{(2)}} \\ & = & I \end{eqnarray} where $\theta = g_{12}^{\rm c} t$. Eq.(\ref{eq:third}) is obtained using Taylor series expansion of the matrix exponents and using $\sigma_x^2 = I$, and Eq.(\ref{eq:fourth}) is obtained using anticommutivity of $\sigma_x$ and $\sigma_z$. The essential features of this decoupling procedure are: \\ \noindent {\bf 1.} The pair of $X^{(2)}$ negates the sign of $\sigma_z^{(2)}$ in the evolution between the pulses. \\ \noindent {\bf 2.} The $X$ pulses make the signs of the $\sigma_z$ matrices of the two spins disagree for exactly half of the time. \\ \noindent {\bf 3.} Since the coupling is bilinear in the $\sigma_z$ matrices, it is unchanged (negated) when the signs of the $\sigma_z$ matrices agree (disagree). The $X$ pulses therefore negate the coupling for exactly half of the time. \\ \noindent {\bf 4.} Since the matrix exponents commute, negating the coupling for exactly half of the time suffices for the evolution to be cancelled out. The most crucial point leading to decoupling is that, the signs of the $\sigma_z$ matrices of the coupled spins, controlled by pairs of $X$ pulses, disagree for half of the time. In general, we consider pulse sequences which concatenate equal-time intervals and use $X$ pulses to control the signs of the $\sigma_z$ of each spin. The essential information on the signs can be represented by a ``sign matrix'' defined as follows. The ``sign matrix'' of a pulse scheme for $n$-spins with $m$ time intervals is the $n \times m$ matrix with the $(i,a)$ entry being the {\em sign} of $\sigma_z^{(i)}$ in the $a$-th time interval. These sign matrices have one-to-one correspondence with our restricted class of pulse sequences. We denote any sign matrix for $n$ spins by $S_n$. For example, the sequence in Eq.(\ref{eq:refocus}) can be represented by the sign matrix \begin{equation} S_2 = \left[ \begin{array}{cc} {+}&{+} \\ {+}&{-} \end{array} \right] \label{eq:s2} \,. \end{equation} The general construction of decoupling scheme is now reduced to finding sign matrices such that every two rows disagree in half of the entries. As a second example, we construct a decoupling scheme for four spins. We first find a correct sign matrix following the previous observations, and then derive the corresponding pulse sequence. For example, a possible sign matrix is given by \begin{equation} S_4 = \left[ \begin{array}{cccc} {+}&{+}&{+}&{+} \\ {+}&{+}&{-}&{-} \\ {+}&{-}&{-}&{+} \\ {+}&{-}&{+}&{-} \end{array} \right] \label{eq:s4} \,, \end{equation} in which any two rows disagree in exactly two entries. $S_4$ can be converted to a pulse scheme by converting each column to a time interval before and after which $X$ pulses are applied to spins (rows) given by $-$'s. No pulses are applied to spins (rows) with $+$'s. The resulting sequence, \begin{eqnarray} \tau (X^{(3)} X^{(4)} \tau X^{(3)} X^{(4)}) (X^{(2)} X^{(3)} \tau X^{(2)} X^{(3)}) \times \nonumber \\ \hfill (X^{(2)} X^{(4)} \tau X^{(2)} X^{(4)}) \label{eq:dec4} \,, \end{eqnarray} is the identity by construction and this can also be verified directly. Note that ${\cal H}_{\rm c}$ in $\tau = e^{-i {\cal H}_{\rm c} t}$ now denotes the sum of six possible coupling terms for four spins. Note also Eq.(\ref{eq:dec4}) is written in such a way that it corresponds visually to the sign matrix, though the evolutions are actually in reverse time order relative to $S_4$. However, such ordering is irrelevant for commuting evolutions. Since $X^{(i)} X^{(i)} = I$, Eq.(\ref{eq:dec4}) can be simplified to \begin{equation} \tau (X^{(3)} X^{(4)} \tau X^{(4)}) (X^{(2)} \tau X^{(3)}) (X^{(4)} \tau X^{(2)} X^{(4)}) \label{eq:dec4short} \,. \end{equation} This simplified pulse sequence can also be obtained directly from Eq.(\ref{eq:s4}) by converting columns to time intervals and inserting $X^{(i)}$ between intervals whenever the $i$-th row changes sign or whenever a $-$ sign reaches either end of the row. Pulse sequences for Eq.(\ref{eq:dec4}) and Eq.(\ref{eq:dec4short}) are shown in Fig.~\ref{fig:pulseseq}. \begin{figure}[ht] \begin{center} \mbox{\psfig{file=s24.eps,width=3.3in}} \vspace*{2ex} \caption{(a) Pulse sequence corresponding to Eq.(\ref{eq:dec4}). From $S_4$, each ``$-$'' sign in the $i$-th row and $a$-th column translates to two $X$ pulses at $\omega_i$ before and after the $a$-th time interval. (b) Pulse sequence obtained from simplifying (a). This corresponds to Eq.(\ref{eq:dec4short}), and can be constructed directly from $S_4$ by translating each change of sign in the $i$-th row to an $X$ pulse at $\omega_i$. A ``$-$'' sign at the end of the row also gives rise to an $X$ pulse at end of the last time interval.} \label{fig:pulseseq} \end{center} \end{figure} The above scheme can be generalized to decouple $n$ spins with $m$ time intervals as follows: \begin{quote} Construct the $n \times m$ sign matrix $S_n$, with entries $+$ or $-$, such that {\em any} two rows disagree in exacly half of the entries. For each $-$ sign in the $i$-th row and the $a$-th column, apply $X^{(i)}$ before and after the $a$-th time interval. \end{quote} Because of the pulses, the sign of the $\sigma_z$ matrix for each spin in each time interval is as given by the sign matrix. The $\sigma_z$ matrices of any two spins therefore have opposite signs for half of the time, so that their coupling is negated for exactly half of the time, and the evolution is always cancelled. For $n$ spins, $n \times m$ sign matrices which correspond to decoupling schemes do not necessarily exist for arbitrary $m$, but they always exist for large and special values of $m$. A possible structure is: \begin{eqnarray} \nonumber S_n = \left[ \begin{array}{cccccccccccc} {+}&{\cdots}&{+}&{+}&{\cdots}&{+}&{+}&{\cdots}&{+}&{+}&{\cdots}&{+} \\ {+}&{\cdots}&{+}&{+}&{\cdots}&{+}&{-}&{\cdots}&{-}&{-}&{\cdots}&{+} \\ { }&{\cdots}&{ }&{ }&{\cdots}&{ }&{ }&{\cdots}&{ }&{ }&{\cdots}&{ } \\ {+}&{\cdots}&{+}&{-}&{\cdots}&{-}&{+}&{\cdots}&{+}&{-}&{\cdots}&{-} \\ { }&{\cdots}&{ }&{ }&{\cdots}&{ }&{ }&{\cdots}&{ }&{ }&{\cdots}&{ } \\ {+}&{\cdots}&{-}&{+}&{\cdots}&{-}&{+}&{\cdots}&{-}&{+}&{\cdots}&{-} \end{array} \right] \,, \end{eqnarray} in which intervals are bifurcated when rows (spins) are added. Such bifurcation takes place whenever it is impossible to add an extra row that is orthogonal to all the existing ones (``depletion''). If such depletion occurs frequently, the sign matrix will have exponential number of columns, and decoupling will take exponential number of steps as $n$ increases. The challenge is to find correct sign matrices with subexponential number of columns. This difficult problem turns out to have solutions given by the Hadamard matrices, which will be decribed next. \subsection{Hadamard matrices} \label{sec:Hadamard} Hadamard matrices have applications in many areas such as the construction of designs, error correcting codes and Hadamard transformations~\cite{CRC96,Sloanehp,vanLint92,MacWilliams77}. A Hadamard matrix of order $n$, denoted by $H(n)$, is an $n \times n$ matrix with entries $\pm 1$, such that \begin{equation} H(n)H(n)^{T} = n I \label{eq:ortho} \,. \end{equation} The rows are pairwise orthogonal, therefore any two rows agree in exactly half of the entries. Likewise columns are pariwise orthogonal. We abbreviate ``$\pm 1$'' as ``$\pm$''. $S_2$ and $S_4$ in Eqs.(\ref{eq:s2}) and (\ref{eq:s4}) are simple examples of $H(2)$ and $H(4)$. An example of $H(12)$ is given by \begin{eqnarray} \nonumber H(12) = \left[ \begin{array}{cccccccccccc} {+}&{+}&{+}&{+}&{+}&{+}& {-}&{+}&{+}&{+}&{+}&{+} \\ {+}&{+}&{+}&{-}&{-}&{+}& {+}&{-}&{+}&{-}&{-}&{+} \\ {+}&{+}&{+}&{+}&{-}&{-}& {+}&{+}&{-}&{+}&{-}&{-} \\ {+}&{-}&{+}&{+}&{+}&{-}& {+}&{-}&{+}&{-}&{+}&{-} \\ {+}&{-}&{-}&{+}&{+}&{+}& {+}&{-}&{-}&{+}&{-}&{+} \\ {+}&{+}&{-}&{-}&{+}&{+}& {+}&{+}&{-}&{-}&{+}&{-} \\ {-}&{+}&{+}&{+}&{+}&{+}& {-}&{-}&{-}&{-}&{-}&{-} \\ {+}&{-}&{+}&{-}&{-}&{+}& {-}&{-}&{-}&{+}&{+}&{-} \\ {+}&{+}&{-}&{+}&{-}&{-}& {-}&{-}&{-}&{-}&{+}&{+} \\ {+}&{-}&{+}&{-}&{+}&{-}& {-}&{+}&{-}&{-}&{-}&{+} \\ {+}&{-}&{-}&{+}&{-}&{+}& {-}&{+}&{+}&{-}&{-}&{-} \\ {+}&{+}&{-}&{-}&{+}&{-}& {-}&{-}&{+}&{+}&{-}&{-} \end{array} \right] \,. \end{eqnarray} The following is a list of useful facts about Hadamard matrices (details and proofs omitted): \begin{enumerate} \item {\em Equivalence}~~Permutations or negations of rows or columns of Hadamard matrices leave the orthogonality condition invariant. Two Hadamard matrices are equivalent if one can be transformed to the other by a series of such operations. Each Hadamard matrix is equivalent to a {\em normalized} one, which has only $+$'s in the first row and column. For instance, $H(12)$ shown previously can be {\em normalized} by negating the 7-${\rm th}$ row and column. \item {\em Necessary conditions}~~$H(n)$ exists only for $n = 1$, $n = 2$ or $n \equiv 0 \bmod 4$. This is obvious if the matrix is normalized, and the columns are permuted so that the first three rows become: \begin{eqnarray} \nonumber \left[ \begin{array}{cccccccccccc} {+}&{\cdots}&{+}&{+}&{\cdots}&{+}& {+}&{\cdots}&{+}&{+}&{\cdots}&{+} \\ {+}&{\cdots}&{+}&{+}&{\cdots}&{+}& {-}&{\cdots}&{-}&{-}&{\cdots}&{-} \\ {+}&{\cdots}&{+}&{-}&{\cdots}&{-}& {+}&{\cdots}&{+}&{-}&{\cdots}&{-} \\ {\cdot}&{\cdot}&{\cdot}&{\cdot}& {\cdot}&{\cdot}&{\cdot}&{\cdot} &{\cdot}&{\cdot}&{\cdot}&{\cdot} \end{array} \right] \,. \end{eqnarray} \item {\em Hadamard's conjecture}~\cite{Hadamard93}~~$H(n)$ exists for every $n \equiv 0 \bmod 4$. This famous conjecture is verified for all $n < 428$. \item {\em Sylvester's construction}~\cite{Sylvester67}~~If $H(n)$ and $H(m)$ exist, then $H(nm)$ can be constructed as $H(n) \otimes H(m)$. In particular, $H(2^r)$ can be constructed as $H(2)^{\otimes r}$, which is proportional to the matrix representation of the Hadamard transformation for $r$ qubits. \item {\em Paley's construction}~\cite{Paley33}~~Let $q$ be an odd prime power. If $q \equiv 3 \bmod 4$, then $H(q+1)$ exists; if $q \equiv 1 \bmod 4$, then $H(2(q+1))$ exists. \item {\em Numerical facts}~\cite{CRC96}~~ For an arbitrary integer $n$, let $\underline{n}$ and $\overline{n}$ be the largest and smallest integers that satisfy $\underline{n} < n \leq \overline{n}$ with {\em known} $H(\overline{n})$ and $H(\underline{n})$. We define the ``gap'', $\delta_n$, to be $\overline{n} - \underline{n}$ (see Fig.~\ref{fig:gap}). For $n\leq 1000$, $H(n)$ is known for every possible order except for 6 cases, and the maximum gap is 8. For $n \leq 10000$, $H(n)$ is unknown for 192 possible orders and the maximum gap is 32. \begin{figure}[ht] \begin{center} \mbox{\psfig{file=gap.eps,width=1.4in}} \vspace*{2ex} \caption{The gap $\delta_n$ between two existing orders of Hadamard matrices.} \label{fig:gap} \end{center} \end{figure} \end{enumerate} The importance of the full connection to Hadamard matrices will become clear after we construct the scheme for an arbitrary number of spins in the next section. \subsection{General Construction of Scheme} In this section, a decoupling scheme for an arbitrary number of spins is constructed. Variations of the scheme to remove Zeeman evolution and to implement selective recoupling are also constructed. The requirements of the scheme are discussed. Recall that $\overline{n}$ is the minimum integer no smaller than $n$ with known $H(\overline{n})$, and $m$ is the number of columns in a sign matrix. For any variations of the scheme for $n$ spins, $m_n$ is defined to be the minimum number of columns in any valid sign matrix and it represents the number of time intervals needed for the intended operation. To construct a decoupling scheme for $n$ spins when $H(n)$ does not necessarily exist, any $H(\overline{n})$ with $\overline{n}-n$ rows omitted can be used as the sign matrix since subsets of rows of $H(\overline{n})$ are still pairwise orthogonal. In other words, any $n \times \overline{n}$ submatrix of $H(\overline{n})$ is a valid sign matrix for decoupling $n$ spins, and $m_n = \overline{n}$. To remove both ${\cal H}_{\rm Z}$ and ${\cal H}_{\rm c}$, we use the fact that the Zeeman term for the $i$-th spin is linear in $\sigma_z^{(i)}$, and negating $\sigma_z^{(i)}$ for half of the time results in no net Zeeman evolution for the $i$-th spin. Therefore, Zeeman evolution for all spins can be removed if the sign matrix has identically zero row sum. Such a sign matrix can be constructed by starting with a {\em normalized} $H(\overline{n})$ and excluding the first row of $H(\overline{n})$ in the sign matrix. Since a normalized $H(\overline{n})$ has only $+$'s in the first row, all other rows have zero row sums by orthogonality. Such construction is possible unless $n = \overline{n}$, in which case construction should start with $H(\overline{n+1})$. Therefore, $m_n = \overline{n}$ if $n < \overline{n}$ and $m_n = \overline{n+1}$ otherwise. To implement selective recoupling between the $i$-th and the $j$-th spins, the sign matrix should have equal $i$-th and $j$-th rows but any other two rows should be orthogonal. The coupling term $g_{ij}^{\rm c} \sigma_z^{(i)} \otimes \sigma_{\rm z}^{(j)}$ never changes sign and that coupling is implemented, while all other couplings are removed. The sign matrix can be obtained from a normalized $H(\overline{n})$ by replacing the $1$-st row with the $j$-th row, and replacing the $j$-th row with the $i$-th row. This scheme also removes ${\cal H}_{\rm Z}$ and $m_n = \overline{n}$. To implement $Z\!\!Z_{ij}$, the duration of each interval $t$ is chosen to satisfy $g_{ij}^{\rm c} \overline{n} t = \pi/4$. Note that the total time used to implement $Z\!\!Z_{ij}$ is the shortest possible, since the coupling is always ``on''. The scheme requires $m_n$ time intervals. Consequently, it requires at most $n m_n$ pulses, since $XX=I$ and the $X$ pulses are only used in pairs. The remaining question is: how does $m_n$ depend on $n$? For simplicity, we consider $\overline{n}$ in place of $m_n$. If Hadamard matrices exist and can be constructed for all orders, $\overline{n} = n$. However, some Hadamard matrices are missing, either because no construction methods are known or they simply cannot exist. Due to missing Hadamard matrices, $\overline{n} = cn$ where $c \geq 1$. We argue in the following that the scheme is still very efficient. First of all, we prove $c < 2$. For each $n$, there exists $r$ such that $2^{r-1} \! \leq n < 2^r$. Since $H(2^r)$ exists by Sylvester's construction, $cn = \overline{n} \leq 2^r \!\! < 2n$. We now show that $c$ is close to the ideal value 1 in most cases, due to the existence of Hadamard matrices of orders other than powers of 2. This is why the full connection to Hadamard matrices is important. In Fig.~\ref{fig:c}, $c$ as a function of $n$ is plotted for $n \leq 10000$. Within this technologically relevant range of $n$, $c$ deviates significantly from 1 only for a few exceptional values of $n$ when $n$ is small. One arrives at the same conclusion by considering the gap $\delta_n = \overline{n} - \underline{n} > \overline{n} - n$, which bounds the extra number of intervals needed in the scheme caused by missing Hadamard matrices. For $n \leq 10000$, $\delta_n/n \ll 1$ except for the few exceptional values of $n$ as a numerical fact. For completeness, we present arguments for $c \approx 1$ for {\em arbitrarily large} $n$ in Appendix~\ref{sec:largen}. This is based on Paley's construction and the prime number theorem. Finally, if Hadamard's conjecture is proven, $\delta_n \leq 3$ $\forall n$. \begin{figure}[ht] \begin{center} \mbox{\psfig{file=cvsnsmall.eps,width=1.7in}\psfig{file=cvsn.eps,width=1.74in}} \vspace*{2ex} \caption{Plots of $c$ vs $n$, where $cn = \overline{n} = m_n$ is the minimun number of time intervals required to perform decoupling or selective recoupling for an $n$-spin system. $c$ for $n \leq 100$ and $101 \geq n \leq 10000$ are plotted separately.} \label{fig:c} \end{center} \end{figure} \section{Conclusion} We reduce the problem of decoupling and selective recoupling in heteronuclear spin systems to finding sign matrices which is further reduced to finding Hadamard matrices. While the most difficult task of constructing Hadamard matrices is not discussed in this paper, solutions already exist in the literature. Even more important is that the connection to Hadamard matrices results in very efficient schemes. Some properties of the scheme are as follows. First of all, the scheme is optimal in the following sense. The rows of Hadamard matrices and their negations form the codewords of first order Reed-Muller codes, which are {\em perfect codes}~\cite{vanLint92,MacWilliams77}. It follows that, for each Hadamard matrix, it is impossible to add an extra row which is orthogonal to all the existing ones. Therefore, for a given $n$, $m_n = \overline{n}$ is in fact the minimum number of time intervals necessary for decoupling or recoupling, if one restricts to the class of pulse sequences considered. Second, the scheme applies for arbitrary duration of the time intervals. This is a consequence of the commutivity of all the terms in the hamiltonian, which in turn comes from the large separations of the Zeeman frequencies compared to the coupling constants. Spin systems can be chosen to satisfy this condition. Finally, disjoint pairs of spins can couple in parallel. We outline possible simplifications of the scheme for systems with restricted range of coupling. For example, a linear spin system with $n$ spins but only $k$-nearest neighbor coupling can be decoupled by a scheme for $k$ spins only. The $i$-th row of the $n \times \overline{k}$ sign matrix can be chosen to be the $r$-th row of $H(\overline{k})$, where $i \equiv r \bmod k$. Selective recoupling can be implemented using a decoupling scheme for $k+1$ spins. The sign matrix is constructed as in decoupling using $H(\overline{k+1})$ but the rows for the spins to be coupled are chosen to be the $k+1$-th row different from all existing rows~\cite{kplus1}. This method involving periodic boundary conditions generalizes to other spatial structures. The size of the scheme depends on $k$ and the exact spatial structure but not on $n$. The scheme has several limitations. First of all, it only applies to systems in which spins can be individually addressed by short pulses and coupling has the simplified form given by Eq.(\ref{eq:dipolar}). These conditions are essential to the simplicity of the scheme. They can all be satisfied if the Zeeman frequencies have large separations. Second, generalizations to include couplings of higher order than bilinear remain to be developed. Furthermore, in practice, RF pulses are inexact and have finite durations, leading to imperfect transformations and residual errors. The present discussion is only one example of a more general issue, that the naturally occuring hamiltonian in a system does not directly give rise to convenient quantum logic gates or other computations such as simulation of quantum systems~\cite{Terhal98}. Efficient conversion of the given system hamiltonian to a useful form is necessary and is an important challenge for future research. \section{Acknowledgments} This work was supported by DARPA under contract DAAG55-97-1-0341 and Nippon Telegraph and Telephone Corporation (NTT). D.L. acknowledges support of an IBM Co-operative Fellowship. We thank Hoi-Fung Chau, Kai-Man Tsang, Hoi-Kwong Lo, Alex Pines, Xinlan Zhou and Lieven Vandersypen for helpful comments.
\section{Introduction} In \cite{stutzki} (hereafter paper I) we defined the $\Delta$-variance as a generalization of the Allan-variance, a method traditionally used in the stability analysis of electronic devices. The Allan- or $\Delta$-variance method allows a clear separation between the global drift characteristics of the signal and other contributions, like white noise. It is specially suited to estimate the spectral index of power law spectral distributions. We demonstrated the use of the two-dimensional $\Delta$-variance for the analysis of molecular cloud images, e.g. maps of CO line emission. Application of the $\Delta$-variance analysis to the integrated CO intensity images of various molecular clouds shows a power law behaviour of power spectrum of the intensity distribution with spectral index $\beta$ close to 2.8. This intensity distribution has a structure well described by a {\em fractional Brownian motion} ({\em fBm}) model and in paper I it was shown that the power law index $\beta$ derived by the $\Delta$-variance is related to the drift exponent $H$ of the corresponding {\em fBm} model by $\beta=E+2H$ (where $E$ is the Euclidean dimension of the model) and to other parameters characterizing the fractal, such as the area-perimeter index. Another tool well established over the last decade in the areas of structure analysis and data processing is the wavelet transform. While the discrete wavelet transform is mainly used in the field of signal compression, the continuous wavelet transform has applications in texture analysis and feature extraction. The connection between the local regularity of a function and the scaling behaviour of its wavelet transform coefficients, as the scale considered becomes smaller and smaller, makes the wavelet technique a natural tool for the analysis of self-similar distributions with (multi-)fractal properties (eg. \cite{holschneider}). In the context of molecular cloud structure analysis the wave\-let transform was first used by \cite{gill}, who also pointed out the advantages of this method over traditional structure indicators like the autocorrelation function. In their study they found a power law behaviour for the centroid velocities of $^{13}$CO spectral data of L1551, as well as for the peak intensities. \cite{langer}, presented a clump decomposition of CO maps of the cloud Barnard 5 using the Laplacian pyramid transform, which is also a wavelet transform. The number distribution of clump masses derived by the Laplacian pyramid transform follows a power law, in agreement with the results presented by \cite{kramer}, obtained by using gaussian clump decomposition. These results give additional evidence for the hierarchical structure of molecular clouds. The purpose of this paper is to show that both these concepts are closely related, namely that the Allan- and $\Delta$-variance can be expressed using the variance of wavelet transform coefficients for suitably chosen wavelets. After a short introduction of the Allan-variance and the wavelet transform in one dimension we show the relation between these concepts (section 2). The use of the $\Delta$-variance and the continuous wavelet transform generalizes our result to two dimensions (section 3). We summarize our results in section 4. \section{Allan-variance and wavelets} \subsection{Allan-variance} We first repeat some facts about the Allan-variance; for a broader introduction the reader may consult appendix A of paper I. To calculate the Allan-variance of a random signal $s(t)$ we consider the differences $d(t,T)$ of succeeding averages over a (time) interval $T$ and calculate their variance. Defining the {\em down-up rectangle function} by \begin{displaymath} \DOWNUP_{T}(t) := \left\{ \begin{array}{r@{\quad }l} \frac{-1}{2T} & \mbox{:\quad} -T \le t < 0 \\ \frac{1}{2T} & \mbox{:\quad} 0 \le t \le T \\ 0 & \mbox{:\quad elsewhere, } \end{array} \right. \end{displaymath} this difference can be written as a convolution \begin{eqnarray} \label{diff} \frac{1}{T} \left( ~ \int\limits_{t-T}^{t} s(t')dt' - \int\limits_{t}^{t+T} s(t')dt' \right) & = & \nonumber\\ -\left\{s * 2 \hspace{.5mm}\DOWNUP_{T}\right\}(t): & = & d(t,T). \end{eqnarray} Without loss of generality we here assume the average $\bar{s}$ to be zero. Then the Allan-variance $\sigma^{2}_{A}(T)$ is defined as \begin{equation} \label{allandef} \sigma^{2}_{A}(T):=\frac{1}{2}\langle d^{2}(t,T) \rangle_{t}, \end{equation} where $\langle ... \rangle_{t}$ denotes the average over all times (or positions) $t$. \subsection{Wavelets} The Fourier transform of a function $\psi$ is defined by $\tilde{\psi}(\omega) = \int_{\Reell} e^{-i 2\pi \omega t} \psi(t) dt$. A square integrable function $\psi$, which satisfies \begin{equation} \label{wavdef} 0 < c_{\psi}:= \int_{\Reell} \frac{| \tilde{\psi}(\omega)|^{2}}{|\omega|} d\omega < \infty \nonumber \end{equation} is called a {\em wavelet}. For $a \in \Reell \backslash \{0\}, b \in \Reell$ the wavelet transform of a signal $s$ with respect to $\psi$ is defined by \begin{equation} \label{wavtrafo} L_{\psi}s(a,b):= \frac{1}{\sqrt{c_{\psi}}} |a|^{-1/2} \int_{\Reell} s(t) \psi\left( \frac{t-b}{a}\right) dt. \end{equation} Each wavelet transform coefficient $L_{\psi}s(a,b)$ is thus the scalar product of the signal with a dilated and translated version of the "mother function" $\psi$. \subsection{The Allan-variance as variance of wavelet transform coefficients} As a special case of a mother function consider a translated and dilated version of the well known Haar wavelet $\psi_{\tilde{H}}$, which we define here as follows: \begin{displaymath} \psi_{\tilde{H}}(t) := \left\{ \begin{array}{r@{\quad }l@{\quad }c} 1 & : & -1 \le t < 0 \\ -1 &: & 0 \le t \le 1 \\ 0 & :& \mbox{elsewhere} \end{array} \right.. \end{displaymath} For this the constant defined by (\ref{wavdef}) is $c_{\psi_{\tilde{H}}}=8 \ln 2$ (see Appendix). In the case of this modified Haar wavelet the transform coefficient $L_{\psi}s(a,b)$ given by (\ref{wavtrafo}) corresponds to the difference of succeeding integrals over $s(t)$, which is the same as the averaging process leading to the definition of the Allan-variance. It follows that \begin{displaymath} \int\limits_{\Reell} \psi_{\tilde{H}}\left( \frac{t-b}{a}\right) s(t) dt= \int\limits_{b-a}^{b} s(t) dt - \int\limits_{b}^{b+a} s(t) dt, \end{displaymath} which equals the term in brackets in eq. (\ref{diff}) with $b \equiv t$ and $a \equiv T$. Further we have \begin{eqnarray} L_{\psi_{\tilde{H}}}s(a,b) & = & \frac{1}{\sqrt{c_{\psi_{\tilde{H}}}}} |a|^{-1/2} \int_{\Reell} s(t) \psi\left( \frac{t-b}{a}\right) dt \nonumber\\ & = & \frac{1}{\sqrt{c_{\psi_{\tilde{H}}}}} |a|^{-1/2} \left( ~ \int\limits_{b-a}^{b} s(t) dt - \int\limits_{b}^{b+a} s(t) dt \right) \nonumber\\ &\stackrel{\mbox{\tiny(\ref{diff})}}{=}& \sqrt{\frac{|a|}{c_{\psi_{\tilde{H}}}}} \cdot d(b,a) \nonumber \\ \mbox{i.e.~~} d(b,a) & = & \sqrt{\frac{c_{\psi_{\tilde{H}}}}{|a|}} L_{\psi_{\tilde{H}}}s(a,b). \end{eqnarray} Using eq. (\ref{allandef}) it follows that the Allan-variance can be expressed as \begin{equation} \label{allanwavelet} \sigma_{A}^{2}(a)=\frac{1}{2}\frac{c_{\psi_{\tilde{H}}}}{|a|} \langle L_{\psi_{\tilde{H}}}^{2}s(a,b) \rangle_{b}, \end{equation} that is the Allan-variance at lag $a$ equals the average of the wavelet transform coefficients over the translation index $b$ for constant dilation index $a$. \section{$\Delta$-Variance and $\Delta$-Wavelet} The main idea of the proceeding section was, that the difference of adjacent averages corresponds with the convolution product of the signal $s$ with an oscillating function $\psi$ of zero mean, that is a wavelet transform coefficient $L_\psi s$. This concept will now be extended to the symmetric version of the Allan-variance, the $\Delta$-variance as introduced in paper I. We recall from paper I (Appendix A, p. 713 ff.) the definition of the one-dimensional {\em down-up-down rectangle function} \begin{displaymath} \DOWNUPDOWN(x) := \left\{ \begin{array}{r@{\quad }l} 1 & \mbox{:\quad} |x| \le 1/2 \\ -1/2 & \mbox{:\quad} 1/2 < |x| \le 3/2 \\ 0 & \mbox{:\quad elsewhere } \end{array} \right. . \end{displaymath} The {\em double difference} $\Delta(t,T)$, that is the average over succeeding intervals is defined as follows: \begin{eqnarray} \label{doppdiff} \Delta(t,T): & = & \left\{s(...)*\frac{1}{T} \DOWNUPDOWN(\frac{...}{T})\right\}(t) \nonumber\\ & = & \int\limits_{\Reell} dx \frac{1}{T} \DOWNUPDOWN\left(\frac{t-x}{T}\right) s(x). \end{eqnarray} The $\Delta$-variance is the variance of the {\em double difference} $\Delta(t,T)$: \begin{equation} \label{delta1var} \sigma_{\Delta}^{2}(T):=\frac{1}{2}\langle \Delta^{2}(t,T) \rangle_{t}. \end{equation} To find the relation to a wavelet transform we choose the {\em down-up-down rectangle function} $\DOWNUPDOWN(x)$ as mother function, which defines the family of one-dimensional "$\Delta$-wavelets"\footnote{A scaled version of this $\Delta$-wavelet is used as "French Hat" wavelet in the literature, cf. \cite{gill}.}: \begin{displaymath} \psi_{\Delta_{a,b}}(x):= \DOWNUPDOWN\left(\frac{x-b}{a}\right), \end{displaymath} where again $a \in \Reell/\{0\}$ and $b \in \Reell$. The value of the admissibility constant is $c_{\psi_{\Delta}} \approx 3.37$. The corresponding wavelet transform coefficients are \begin{displaymath} L_{\psi_{\Delta}}s(a,b) = \frac{1}{\sqrt{c_{\psi_{\Delta}}}} |a|^{-1/2} \int\limits_{\Reell} dx \DOWNUPDOWN\left(\frac{x-b}{a}\right) \cdot s(x), \end{displaymath} and with $a \equiv T, b \equiv t$ and $\DOWNUPDOWN(x)=\DOWNUPDOWN(|x|)$ one gets \begin{eqnarray} \Delta(b,a) &\stackrel{\mbox{\tiny(\ref{doppdiff})}}{=} & \sqrt{\frac{c_{\psi_{\Delta}}}{|a|}} L_{\psi_{\Delta}}s(a,b). \nonumber \end{eqnarray} With eq. (\ref{delta1var}) this leads to the result analogous to eq. (\ref{allanwavelet}): \begin{equation} \label{deltawavelet} \sigma_{\Delta}^{2}(a)=\frac{1}{2} \frac{c_{\psi_{\Delta}}}{|a|}\langle L^{2}_{\psi_{\Delta}}s(a,b)\rangle_{b}. \end{equation} \subsection{Two-dimensional Continuous Wavelet Transform} To generalize our results to higher dimensions we make use of the definition of two-dimensional (directional) wavelets given by \cite{anto}. As in the one-dimensional case a two-dimensional function $\psi$ is called a wavelet, if the following admissibility condition holds: \begin{displaymath} 0 < c_{\psi}:= \int_{\Reell^{2}} \frac{| \tilde{\psi}(\vec{f})|^{2}}{\|\vec{f}\|^{2}} d^{2}\vec{f} < \infty. \end{displaymath} From \cite{anto} we adopt the notation for translation, dilation and rotation operators:\\ \vspace*{2mm} \\ $\mbox{translation: \quad} (T^{\vec{b}}s)(\vec{x}) = s(\vec{x}-\vec{b}), ~~b\in \Reell^{2} \\ \mbox{dilation: \hspace{0.35cm} \quad} (D^{a}s)(\vec{x}) = \frac{1}{a}s(\frac{\vec{x}}{a}), ~~a>0 \\ \mbox{rotation:\hspace{0.4cm} \quad} (R^{\theta}s)(\vec{x}) = s(r_{-\theta} (\vec{x})), ~~\theta\in[0,2\pi),\\$\vspace*{2mm}\\ where $r\in SO(2)$. A combination $\Omega(a,\theta,\vec{b}) = T^{\vec{b}}D^{a}R^{\theta}$ of these operators acts on a function $s$ in the following way: \begin{displaymath} (\Omega(a,\theta,\vec{b})s)(\vec{x})=s_{a,\theta,\vec{b}}(\vec{x}) \equiv \frac{1}{a} s\left(\frac{1}{a} r_{-\theta} (\vec{x}-\vec{b})\right). \end{displaymath} With this notation the two-dimensional continuous wave\-let transform (CWT) of a function (or ``image'') $s$ with respect to $\psi$ is given by \begin{equation} \label{wav2trafo} L_{\psi}s(a,\theta,\vec{b}):= \frac{|a|^{-1}}{\sqrt{c_{\psi}}} \int_{\Reell^{2}} d^{2}\vec{x}~ \psi\left(r_{-\theta}(\frac{\vec{x}-\vec{b}}{a})\right) s(\vec{x}) \end{equation} with $a \in \Reell \backslash \{0\}, \vec{b} \in \Reell^{2}$. For a complex valued mother function $\psi$ in eq. (\ref{wav2trafo}) is replaced by its complex conjugate $\psi^{*}$, but we will only make use of real valued functions here. At present only isotropic wavelets $\psi(\vec{r})=\psi(r), r=|\vec{r}|$ will be considered, so that we omit the index $\theta$: \begin{displaymath} L_{\psi}s(a,\vec{b}):= \frac{|a|^{-1}}{\sqrt{c_{\psi}}} \int_{\Reell^{2}} d^{2}\vec{x}~ \psi\left( \frac{\vec{x}-\vec{b}}{a}\right) s(\vec{x}). \end{displaymath} \subsection{$\Delta$-Variance in Two Dimensions} This section is based on paper I, Appendix B (p. 717), where we now consider the case $E=2$, that is two-dimensional signals or images. The two-dimensional {\em down-up-down rectangle function} is given by: \begin{eqnarray} \label{twodmother} \stackrel{2}{\DOWNUPDOWN}(\vec{r} ) & : = & \frac{3^{2}}{(3^{2}-1)}\left[~\sqcap(\vec{r}) - \frac{1}{3^{2}}\sqcap(\frac{\vec{r}}{3})~\right] \nonumber\\ & = & \left\{ \begin{array}{r@{\quad }l@{\quad }c} 1 & : & |\vec{r}| \le 1/2 \nonumber\\ -1/8 & : & 1/2 < |\vec{r}| \le 3/2 \\ 0 & : & \mbox{ elsewhere } \end{array} \right.\nonumber. \end{eqnarray} For the two-dimensional unit sphere the "volume" is ${\cal V}_{2}= \pi$ and the "surface" is ${\cal S}_{2}=2 \pi$. Using the nomenclature of paper I we have \begin{equation} \label{twodfilter} \bigodot_{D,2}(\vec{r})=\frac{1}{\pi (\frac{D}{2})^{2}} \stackrel{2}{\DOWNUPDOWN}(\frac{\vec{r}}{D}). \end{equation} Convolving $s(\vec{r})$ with $\bigodot_{D,2}(\vec{r})$ leads to the difference of averages at average distance $D$: \begin{eqnarray} \label{diff2} \Delta_{s}(\vec{r},D,2) & := & \left\{ s(...) * \bigodot_{D,2}(...)\right\}(\vec{r}) \nonumber\\ & = & \int\limits_{\Reell^{2}} d^{2}\vec{x} \bigodot_{D,2}(\vec{r}-\vec{x}) \cdot s(\vec{x}) \nonumber\\ & = & \frac{1}{\pi \left(\frac{D}{2}\right)^{2}} \int\limits_{\Reell^{2}} d^{2}\vec{x} \hspace{.8mm} \DOWNUPDOWN\left(\frac{\vec{r}-\vec{x}}{D}\right) \cdot s(\vec{x}) \end{eqnarray} As in the one-dimensional case the $\Delta$-variance is given as the variance, i.e the autocorrelation function at zero lag of the filtered signal: \begin{equation} \label{sigma2dim} \sigma^{2}_{\Delta_{s}}(D):=\frac{1}{2 \pi} A_{\Delta_{s}(\vec{r},D,2)}(0). \end{equation} \subsection{$\Delta$-Variance and Two-dimensional Continuous Wave\-let Transform} To relate the $\Delta$-variance to the two-dimensional CWT one naturally considers the {\em down-up-down-rectangle function} given by eq. (\ref{twodmother}) as mother wavelet and defines \begin{displaymath} \psi_{\Delta_{a,\vec{b}}}(\vec{r}):= \stackrel{2}{\DOWNUPDOWN}\left(\frac{\vec{r}-\vec{b}}{a} \right); \end{displaymath} $\psi_{\Delta_{a,\vec{b}}}$ will also be called (two-dimensional) {\em $\Delta$-wavelet}; the admissibility constant for the two-dimensional function is $c_{{\psi}_{\Delta}} \approx 4.11$. Like in the one-dimensional case the wavelet transform coefficients $L_{\psi_{\Delta}}s(a,\vec{b})$ are directly related to the differences of average values at average distance $D$ $\Delta(\vec{r},D,2)$ (cf. eq. (\ref{diff2})): \begin{eqnarray} \label{wavdelta} L_{\psi_{\Delta}}s(a,\vec{b}) & = & \frac{1}{\sqrt{c_{\psi_{\Delta}}}} \frac{1}{a} \int\limits_{\Reell^{2}} d^{2}\vec{x} \hspace{.8mm} \psi_{{\Delta}_{a,\vec{b}}}(\vec{x}) \cdot s(\vec{x}) \nonumber\\ & = & \frac{1}{\sqrt{c_{\psi_{\Delta}}}} \frac{1}{a} \int\limits_{\Reell^{2}} d^{2}\vec{x} \hspace{.8mm} \DOWNUPDOWN \left(\frac{\vec{x}-\vec{b}}{a}\right) \cdot s(\vec{x}) \end{eqnarray} We identify $D \equiv a$ and $\vec{r} \equiv \vec{b}$ and because of $\DOWNUPDOWN(\vec{r})=\DOWNUPDOWN(|\vec{r}|)$ we find using eq. (\ref{diff2}) and eq. (\ref{wavdelta}): \begin{eqnarray} L_{\psi_{\Delta}}s(a,\vec{b}) & = & \frac{1}{\sqrt{c_{\psi_{\Delta}}}} \frac{1}{a} \cdot \frac{\pi a^{2}}{4} \cdot \Delta_{s}(\vec{b},a,2) \nonumber\\ & = & \frac{\pi a}{4 \cdot \sqrt{c_{\psi_{\Delta}}} } \cdot \Delta_{s}(\vec{b},a,2). \nonumber \end{eqnarray} Looking at the autocorrelation function as the two-dimensional generalization of the variance as we have done in eq. (\ref{sigma2dim}), we find as the main result of this section: \begin{equation} \label{mainresult} \sigma^{2}_{\Delta_{s}}(a)= \frac{1}{2 \pi} \frac{16 \hspace{.8mm} c_{\psi_{\Delta}}}{\pi^{2} a^{2}} \hspace{.8mm} A_{L_{\psi_{\Delta}}s(a,\vec{b})}(0), \end{equation} i.e. the $\Delta$-variance for a given lag $a$ can be expressed in terms of the autocorrelation function at zero lag of the wavelet transform coefficients with dilation parameter $a$. \section{Summary and Outlook} The main results of our study are equations (\ref{allanwavelet}), (\ref{deltawavelet}) and (\ref{mainresult}) which show, that the Allan- and $\Delta$-variance can be expressed using the coefficients of special wavelet transforms and that both these concepts are very closely related. \cite{pando} have shown that the {\em discrete} wavelet transform is a good power spectrum estimator especially in the case of finite sized samples. The same result was found by \cite{bensch} for the $\Delta$-variance analysis of CO integrated intensity maps as small as $32 \times 32$ pixel. Detailed investigations by Bensch et al. and \cite{ossk} show that the $\Delta$-variance remains a robust measure of astronomical images if the effects of noise, smearing due to the finite size antenna beam pattern and radiative transfer are taken into account. Pando et al. pointed out that evaluating the power spectrum with an over-complete basis function system of the CWT can lead to correlations "that are not in the sample, but due to correlations among the wavelet coefficients". This will also be true for the Allan- and the $\Delta$-variance, because the definitions of both employ such an over-complete system. It has to be checked in a further study, how the results change, if only coefficients on a discrete, wavelet adapted grid (a "wavelet frame") are considered, i.e. if the continuous wavelet transform is replaced by a discrete one. While the choice of isotropic wavelets was natural and sufficient in our work on the isotropic scaling behaviour of molecular cloud structure, the directional sensitivity of non-axial symmetric wavelets will allow the determination of preferred orientations, eg. of filamentary substructures, in astronomical images. \begin{appendix} \section{Evaluation of the factors $c_{\psi}$} For all wavelets discussed in this paper the factors \begin{displaymath} c_{\psi}:= \int_{\Reell} \frac{| \tilde{\psi}(\omega)|^{2}}{|\omega|} d\omega \mbox{\quad and \quad} c_{\psi}:= \int_{\Reell^{2}} \frac{\| \tilde{\psi}(\vec{f})\|^{2}}{\|\vec{f}\|^{2}} d^{2}\vec{f} \end{displaymath} for the one- and two-dimensional case, respectively, can be calculated analytically. The integrals were solved using {\em Mathematica 3.0 for students}. \subsection{Modified Haar-Wavelet} Looking at the mother function $\psi(x)=-2 \DOWNUP(x)$ of the modified Haar wavelet system we find (paper I, p.714) \begin{displaymath} | \tilde{\psi}(\omega)|^{2}=4 \frac{\sin^{4}(\pi \omega)}{(\pi \omega)^{2}} \mbox{\quad and therefore} \end{displaymath} \begin{eqnarray} c_{\psi_{\tilde{H}}} & = & \int\limits_{- \infty}^{\infty} 4 \frac{\sin^{4}(\pi \omega)}{(\pi \omega)^{2} \cdot |\omega|} \nonumber = \int\limits_{- \infty}^{\infty} 4\pi \frac{\sin^{4}(\pi \omega)}{|\pi \omega|^{3}}\nonumber = 8\ln 2 \nonumber \end{eqnarray} \subsection{$\Delta$-Wavelet} \subsubsection{$\Delta$-Wavelet in One Dimension} Looking at the mother function $\psi(x)=\DOWNUPDOWN(x)$ we find (paper I, p.715) \begin{displaymath} | \tilde{\psi}(\omega)|^{2}=4 \frac{\sin^{6}(\pi \omega)}{(\pi \omega)^{2}} \mbox{\quad and therefore} \end{displaymath} \begin{eqnarray} c_{\psi_{\Delta}} & = & \int\limits_{- \infty}^{\infty} 4 \frac{\sin^{6}(\pi \omega)}{(\pi \omega)^{2} \cdot |\omega|}\nonumber = \int\limits_{- \infty}^{\infty} 4\pi \frac{\sin^{6}(\pi \omega)}{|\pi \omega|^{3}} \nonumber\\ & = & 12\ln 2 - \frac{9 \ln 3}{2} \approx 3.37 \nonumber \end{eqnarray} \subsubsection{$\Delta$-Wavelet in Two Dimensions} The power spectrum of the two-dimensional {\em down-up-down rectangle function} is (paper I, p. 717) \begin{displaymath} \|\widetilde{\DOWNUPDOWN}(\vec{f})\|^{2}=\left[ \frac{J_{1}(\pi f)}{(\pi f)} - \frac{J_{1}(3 \pi f)}{(3 \pi f)}\right]^{2}, \end{displaymath} where $J_{1}$ denotes the first order {\em Bessel}-function of first kind. With this we get \begin{eqnarray} c_{\psi_{\Delta}} & = & \int\limits_{\Reell^{2}} \frac{\| \widetilde{\DOWNUPDOWN}(\vec{f})\|^{2}}{\|\vec{f}\|^{2}} d^{2}\vec{f = 2\pi \int\limits_{0}^{\infty} \frac{\left[ \frac{J_{1}(\pi f)}{(\pi f)} - \frac{J_{1}(3 \pi f)}{(3 \pi f)}\right]^{2}}{f} df \nonumber\\ & = & 2\pi \left(\ln(3)-\frac{4}{9}\right) \approx 4.11 \nonumber \end{eqnarray} \end{appendix}
\section{Introduction} There are many natural phenomena where the dynamics does not take place smoothly but in the form of bursts or avalanches. Avalanche dynamics has been observed in earthquakes dynamics \cite{richter}, granular piles \cite{held,frette1}, superconductor vortex piles \cite{superconductors}, the Barkhausen effect \cite{barkhausen}, crack propagation \cite{crack}, and more. The avalanches are characterized by their size $s$ and duration $T$ which in general follow the power law distributions $P(s)\sim s^{-\tau_s}$ and $P(T)\sim T^{-\tau_t}$, respectively. Bak, Tang, and Wiesenfeld \cite{bak2} introduced the notion of self-organized criticality (SOC) to explain this kind of dynamics, which seems to be common to many "different" phenomena. In their early formulation the critical state of a SOC system is an attractor of its dynamics and therefore there is no need of fine-tuning to reach the critical state. This idea was illustrated with very simple cellular automata, the sandpile models. In sandpile models an integer \cite{bak2,kadanoff,manna1,frette,nunes1} or continuous \cite{zhang,olami,bassler} variable $z_i$ is defined in a $d$-dimensional lattice. $z_i$ can be the height of grains columns \cite{bak2,kadanoff,nunes1,christensen}, the stress accumulated in certain fault \cite{olami}, the vortex density \cite{bassler}, etc. A critical threshold condition for local relaxation (toppling) is also given, which can be a critical height $z_i>z_c$ \cite{bak2,kadanoff,manna1,olami}, critical slope $m_i=z_i-z_{i+1}>m_c$ \cite{kadanoff,frette,bassler}, or critical Laplacian $l_i=\sum_{nn}z_i-2dz_i>l_c$ \cite{manna2} threshold condition. When this condition is fulfilled the site relaxes transferring grains to its nearest neighbors (nn), otherwise certain amount of grains is added from an external field. Critical height models have been extensively studied in the literature, either by mean field \cite{MF} and field theories \cite{vespignani1}, renormalization group (RG) \cite{diaz-guilera,pietronero}, and numerical simulations \cite{bak2,kadanoff,manna1,zhang,olami}. However, the study of critical slope models has been in general limited to numerical simulations \cite{kadanoff,frette,bassler}, which are not so extensive as in the critical height variant, while theoretical approaches are almost absent. There are many experimental situations, sandpiles and vortexpiles for instance, where the critical slope threshold condition is more appropriate, requiring a rigorous treatment which captures the general features of critical slope models. In this direction Paczuski and Boettcher \cite{paczuski} provided an starting point. They mapped a critical slope model into a linear interface depinning (LID) model where the interface is dragged at one end, showing the existence of common behavior between these models. The importance of their observation is that LID models can be solved using continuous approaches \cite{nattermann,narayan1,narayan2}. Their work was inconclusive because, up to our knowledge, the problem of an interface moving on a random environment and dragged at one end has not been solved. However, they observed that some scaling exponents, but not all, are identical to those measured for an interface driven uniformly. In the present work we solve the problem of an interface moving on a random environment and dragged at one end. We start with a LID equation with the appropriate boundary conditions. We then look for the fluctuations around the average in the stationary state. In the thermodynamic limit these fluctuations satisfy an equation similar to that of an interface driven uniformly. The critical exponents are then computed using previous RG calculations for the uniformly driven interface \cite{nattermann,narayan1,narayan2}. The dynamic and roughness scaling exponents are found identical while other exponents become different. These results show a very good agreement when compared with numerical estimates of the scaling exponents reported in the literature. \section{The model} Consider the following one-dimensional critical slope model. An integer variable $z_i$ ($i=1,2,\ldots,L$) is defined. A site where the slope $m_i=z_i-z_{i+1}$ is greater than a threshold $m_c$ is said to be active and relaxes according to the toppling rule \begin{eqnarray} z_i\rightarrow z_i-1, \nonumber\\ z_{i+1}\rightarrow z_{i+1}+1. \label{eq:0} \end{eqnarray} Grains are added at site $i=0$ at rate $c$ while the other boundary $i=L-1$ is open. Usually $c$ is assumed very small in such a way that a new grain is added only when no site is active, which is the usual separation of time scales assumed in numerical simulations. Paczuski and Boettcher \cite{paczuski} noted that instead of follow the evolution of the column heights $z_i$ one can follow the evolution of $h_i(t)$, the total number of toppling events at site $i$ up to time $t$. The evolution rules for $h_i$ can be easily obtained from those for $z_i$. If we start with an initial configuration where $z_i=0$ the number of grains at site $i$ is obtained adding the number of grains received from the left neighbor ($h_{i-1}$) and subtracting the number of grains transferred to the right neighbor ($h_i$), i.e. \begin{equation} z_i=h_{i-1}-h_i. \label{eq:0a} \end{equation} Using this expression one may compute the local slope $m_i=z_i-z_{i+1}$, resulting \begin{equation} m_i-m_c=h_{i+1}+h_{i-1}-2h_i-m_c, \label{eq:1} \end{equation} On the other hand, $h_i$ increases by one unit every time the site $i$ topples, i.e. every time $m_i>m_c$, which can be written as \begin{equation} h_i(t+1)-h_i(t)=\Theta(m_i-m_c), \label{eq:1a} \end{equation} where $\Theta(x)$ is the Heaviside unit step function. The evolution rules in eqs. (\ref{eq:1}) and (\ref{eq:1a}) are just the discretized evolution rules for LID models \cite{leschhorn}, taking $m_i-m_c$ as the force acting on the interface. The left term of eq. (\ref{eq:1a}) is a discrete time derivative while $h_{i+1}+h_{i-1}-2h_i$ is the discrete Laplacian in one dimension. After coarse-graining one obtains the continuum equation \cite{leschhorn} \begin{equation} \lambda \partial_t h=\Gamma\nabla^2h-m_c+\eta(x,h). \label{eq:2} \end{equation} Here $\lambda$ and $\Gamma$ are coarse-graining parameters, $\lambda$ can be interpreted as a viscosity coefficient and $\Gamma$ as a surface tension. $\eta(x,h)$ is a quenched noise which may have different origins. For instance, the critical slope may be a random variable reflecting the randomness in the local rearrangements of sand grains after each toppling. In a more realistic model $m_c$ is thus a random variable which changes its value from site to site after each toppling event. When $h(x,t)$ advances, i.e. the site $x$ topples, a new random slope is assigned. The randomness in the critical slope is thus reflected in eq. (\ref{eq:2}) through the quenched noise $\eta(x,h)$. The quenched noise $\eta(x,h)$ will be assumed a Gaussian noise with zero mean and noise correlator \begin{equation} \langle \eta(x,h)\eta(x^\prime,t^\prime)\rangle= \delta^d(x-x^\prime)\Delta(t-t^\prime), \label{eq:2aa} \end{equation} where $\Delta(h)$ is a symmetric function, i.e. $\Delta(-h)=\Delta(h)$. Depending on the boundary conditions for $\Delta(h)$ one can distinguish two cases \cite{nattermann,narayan1,narayan2}. For random disorder $\Delta(h)$ goes to zero for large $h$ while for periodic pinning forces $\Delta(h)$ is periodic. The problem will be completely defined after the initial and boundary conditions are specified. We are interested in the stationary solution of eq. (\ref{eq:2}) and therefore the initial condition is irrelevant. Grains will be added to the system at constant rate $c$ at site $i=0$. Hence, if we start with an initial configuration where $z_0=0$ the number of grains at site $i=0$ is obtained adding the number of grains received from the external field ($ct$), and subtracting the number of grains transferred to the right neighbor ($h_i$), i.e. \begin{equation} z_i=ct-h_i. \label{eq:0aa} \end{equation} Using this expression one may compute the local slope $m_i=z_i-z_{i+1}$, resulting \begin{equation} m_i-m_c=h_1-2h_0-m_c. \label{eq:0b} \end{equation} This evolution rule together with that in eq. (\ref{eq:1a}) leads to the coarse grained equation \begin{equation} \lambda \partial_t h=\kappa\frac{\partial h}{\partial x}+ ct-h-m_c+\eta(x,h), \label{eq:0c} \end{equation} which is the boundary condition at $x=0$. However, in the stationary state the major contribution to the right hand side of this equation is given by the term $ct-h$ and, therefore, this boundary condition can be approximated by the more simple condition \begin{equation} h(0,t)=ct. \label{eq:0d} \end{equation} On the other hand, at $x=L-1$ the boundary is open which implies that the site $x=L$ will never topples, i.e. \begin{equation} h(L,t)=0. \label{eq:0e} \end{equation} Eq. (\ref{eq:2}) can be generalized to a $d$-dimensional model. To be more specific we consider a $d$-dimensional hyper-cube of linear size $L$ where grains are added at face $x_{||}=0$ and the boundary $x_{||}=L$ is open. The equation of motion of $h(x,t)$ will be given by eq. (\ref{eq:2}), where $\nabla^2$ will be now the $d$-dimensional Laplacian, with the boundary conditions \begin{equation} h|_{x_{||}=0}=ct,\ \ \ \ h|_{x_{||}=L}=0. \label{eq:3} \end{equation} Periodic boundary conditions are assumed in the other directions. Eqs. (\ref{eq:2}) and (\ref{eq:3}) completely defines the LID model dragged at one end. For $c\rightarrow0$ we recover the equation of motion proposed by Paczuski and Boettcher \cite{paczuski}, which describes the dynamics of critical slope models in the limit of separation of time scales and no local dissipation. \section{Stationary state} Eq. (\ref{eq:2}) is similar to the equation of motion of LID models. However, in this case the interface is dragged at one end instead of being driven by a constant force. The drag at one end makes the problem asymmetric, which leads to a gradient in $h(x,t)$. It is thus easier to look for an expansion around this gradient. With this idea in mind we look for a solution of the form \begin{equation} h(x,t)=h_0(x_{||},t)+y(x,t), \label{eq:4} \end{equation} where $h_0(x,t)$ will be determined imposing the constraint $y|_{x_{||}=0}=y|_{x_{||}=L}=0$ (symmetric boundary conditions for $y(x,t)$) and introducing a constant force term in the equation for $y(x,t)$. These requirements lead to the following problem for $h_0(x,t)$ \begin{equation} \Gamma \frac{\partial^2 h_0}{\partial x_{||}^2}=F-m_c, \label{eq:4a} \end{equation} with the boundary conditions in eq. (\ref{eq:3}). The solution of this problem is given by \begin{equation} h_0(x,t)=\frac{F-m_c}{2\Gamma}x_{||}^2-\left(ct+ \frac{F-m_c}{2\Gamma}L^2\right)\frac{x_{||}}{L}+ct. \label{eq:4b} \end{equation} Then, substituting eq. (\ref{eq:4}) in eq. (\ref{eq:2}), with $h_0(x,t)$ given by eq. (\ref{eq:4b}), and taking the limits $vt\gg(F-m_c)L^2/2\Gamma$ and $x_{||}\ll L$ we obtain the following equation for $y(x,t)$ \begin{equation} \lambda \partial_t y=\Gamma\nabla^2y+ F-\lambda c+\eta(x,ct+y). \label{eq:7} \end{equation} This equation describe the fluctuations of the interface profile $h(x,t)$ around $h_0(x_{||},t)$, away from the boundaries and for very long times. The constant force $F$, introduced in eq. (\ref{eq:4a}), will be determined self-consistently using the constraint $\langle y(x,t)\rangle=0$. Within this approximations the anisotropy introduced by the boundary conditions, such that the resulting equation for $y(x,t)$ is isotropic. The influence of the anisotropy will be considered in the next subsection to compute the avalanche exponents. Eq. (\ref{eq:7}) is identical to the one obtained for the fluctuations of an elastic interface driven uniformly by a constant force $F$ \cite{nattermann,narayan1}. However, in this case $F$ is a fixed parameter while the interface velocity $v$ ($c$ in eq. (\ref{eq:7})) is obtained self-consistently from the equation of motion. A depinning transition takes place at certain critical force $F_c$ determined by the disorder. For $F<F_c$ the interface is pinned after certain finite time while for $F>F_c$ it moves with finite average velocity $v\sim(F-F_c)^\beta$. On the contrary, in our model the interface velocity $c$ is the fixed parameter, while $F$ is determined self-consistently from the equation of motion. Since $c>0$ the system will always be above the depinning transition, i.e. $F>F_c$. Moreover, to obtain an average interface velocity $c$ we should have $c\sim(F-F_c)^\beta$ and therefore \begin{equation} F=F_c+\text{const.}c^{1/\beta}. \label{eq:8} \end{equation} To reach the critical state $F=F_c$ we must then fine-tune $c$ to zero. In other words, the critical state will be obtained in the limit of separation of time scales $c\rightarrow0$, which is usually satisfied in numerical simulations. According to eq. (\ref{eq:4b}) adjusting the constant force $F$ we are just adjusting the curvature of the interface profile $h(x,t)$ along the $x_{||}$ direction. Hence, the system self-organizes into a stationary state where the curvature of the interface balances the pinning forces. This conclusion, which was already conjectured by Paczuski and Boettcher, was obtaining here by solving the LID model dragged at one end. \subsection{Scaling exponents} The fluctuations around the average profile are characterized by the pair correlation function. At the critical state the pair correlation function is expected to obey the scaling law \cite{nattermann,narayan1} \begin{equation} \langle[y(x,t)-y(0,0)]^2\rangle \sim |x|^{2\zeta}g(t/|x|^z), \label{eq:11} \end{equation} where $g(x)$ is an scaling function and $z$ and $\zeta$ are the dynamic and roughness exponents, respectively. These exponents are obtained through a RG analysis of eq. (\ref{eq:7}) \cite{nattermann,narayan1,narayan2}. The upper critical dimension is found to be $d_c=4$ and below $d_c$ the scaling exponents are given by \begin{equation} z=2-\frac{2}{9}(4-d),\ \ \ \ \zeta=\frac{4-d}{3}, \label{eq:12} \end{equation} for a random distribution of pinning forces \cite{nattermann,narayan1} and \begin{equation} z=2-\frac{4-d}{3},\ \ \ \ \zeta=0, \label{eq:12a} \end{equation} for periodic distribution of pinning \cite{narayan2}. The exponents $z$ and $\zeta$ are thus identical to those obtained for the constant force case. However, other scaling exponents result different because of the anisotropy introduced by the boundary conditions. In the case of an interface driven by a constant force the average interface profile above the depinning transition is given by $vt$ and, therefore, there is no preference direction in space. On the contrary, in the case of an interface dragged at one end the average interface profile is given by $h_0(x_{||},t)$, which is clearly non-uniform along the $x_{||}$ direction as one can see from eq. (\ref{eq:4b}). To analyze the influence of the anisotropy introduced by the boundary conditions let us analyze the dynamics of active sites. Let $\rho_a(x_{||},x_\bot,t)$ the average density of active sites at site $(x_{||},x_\bot)$ and time $t$, given was a site active at $x_{||}=0$. Here $x_{||}$ as above denotes the position along the preferential direction and $x_\bot$ is a $d-1$ dimensional vector denoting the position in a plane perpendicular to $x_{||}$. Since the flow of grains takes place, in average, along the positive $x_{||}$ direction then the average flux of particles through the plain $x_{||}=L$ is given by \begin{equation} J(L)=\int dt d^{d-1}x_\bot\rho_a(L,x_\bot,t). \label{eq:12b} \end{equation} On the other hand, at the critical state $c\rightarrow 0$ the average density of active sites should satisfy the scaling law \begin{equation} \rho_a(x_{||},x_\bot,t)=t^{\eta-d/z}f\left(\frac{x_{||}}{t^{1/z}}, \frac{|x_\bot|}{t^{1/z}},\frac{t}{L^z}\right), \label{eq:12c} \end{equation} where $\eta$ is another scaling exponent and $f$ is an scaling function. Then, substituting this expression in eq. (\ref{eq:12b}) it results that \begin{equation} J(L)\sim L^{(1+\eta)z-1}. \label{eq:13c} \end{equation} Now, if the system is in a stationary state for each grain we put at $x_{||}=0$ one grain should go out at $x_{||}=L$ and, therefore, $J(L)=1$. This stationary condition will be satisfied only if \begin{equation} (1+\eta)z=1, \label{eq:14c} \end{equation} in eq. (\ref{eq:13c}. A more familiar scaling relation is obtained if one computes the mean avalanche size, which is given by \begin{equation} \langle s\rangle=\int dt dx_{||}d^{d-1}x_\bot \rho_a(x_{||},x_\bot,t). \label{eq:14d} \end{equation} Substituting the scaling law for $\rho_a$ in eq. (\ref{eq:12c}) in this expression it results that \begin{equation} \langle s\rangle \sim L^{(1+\eta)z}\sim L, \label{eq:14e} \end{equation} which is the usual scaling dependency of the mean avalanche size with system size in critical slope models. This derivation using such simple scaling arguments is reported here for the first time. Using the RG estimates for the exponents $z$ and $\zeta$ in eqs. (\ref{eq:12}) or (\ref{eq:12a}) and the scaling relation $\langle s\rangle\sim L$ we can compute the avalanche scaling exponents. Let $s$ be the avalanche size and $T$ its duration, which are distributed according to $P(s)$ and $P(T)$, respectively. Just at the critical state one expect that these distributions satisfy the power law behavior $P(s)\sim s^{-\tau_s}$ and $P(T)\sim T^{-\tau_t}$, where $\tau_s$ and $\tau_t$ are the avalanche distribution exponents. However, for a system of finite size a characteristic avalanche size $s_c\sim L^D$ and duration $T_c\sim L^z$ will appear, where $D$ is the avalanche fractal dimension. Near the critical state the distributions of avalanche size and duration will thus satisfy the scaling laws \begin{equation} P(s)\sim s^{-\tau}f(s/L^D),\ \ \ \ P(T)\sim T^{-\tau_t}g(T/L^z), \label{eq:13} \end{equation} where $f(x)$ and $g(x)$ are some cutoff functions with the asymptotic behaviors $f(x),g(x)\sim 1$ for $x\ll1$ and $f(x),g(x)\ll1$ for $x\gg1$. The exponents $\tau_s$, $\tau_t$, D and $z$ are not all independent. Since $s\sim T^{z/D}$ then the condition $\int dsP(s)=\int dTP(T)$ implies \begin{equation} (\tau_s-1)D=(\tau_t-1)z. \label{eq:14} \end{equation} Another scaling relation can be obtained taking into account that $\langle s\rangle\sim L$ (see eq. (\ref{eq:14e}), resulting \begin{equation} (2-\tau_s)D=1. \label{eq:15} \end{equation} Then, from eqs. (\ref{eq:14}) and (\ref{eq:15}) it follows that \begin{equation} \tau_s=2-\frac{1}{D},\ \ \ \ \tau_t=1+\frac{D-1}{z}. \label{eq:15a} \end{equation} Finally there is a scaling relation which relates the avalanche dimension exponent $D$ with the roughness exponent $\zeta$. Below the upper critical dimension the avalanches are compact objects \cite{paczuski3} and therefore $s\sim \Delta h r^d$, where $\Delta h$ is the characteristic fluctuation of the interface width during the avalanche and $r$ its characteristic linear extent in the $d$-dimensional substrate. Then since $\Delta h\sim r^\zeta$ and $s\sim r^D$ one obtains \cite{paczuski3} \begin{equation} D=d+\zeta \label{eq:16} \end{equation} Above the upper critical dimension the avalanches are no more compact and $D=d_c=4$ \cite{paczuski3}. \section{Comparison with experiments and numerical simulations} Using the values of $z$ and $\zeta$ obtained from the RG analysis in eq. (\ref{eq:12}) or (\ref{eq:12a}) and the scaling relations in eqs. (\ref{eq:15a}) and (\ref{eq:16}) we can determine all the avalanche exponents. The results in one and two dimensions are shown in table \ref{tab:1} and \ref{tab:2} for random and periodic pinning, respectively. Some numerical estimates for critical slopes model are also shown for comparison. In $d=1$ we count with numerical simulations of ricepile models, which are critical slope sandpile models with certain randomness in the toppling rule. For instance Frette \cite{frette} considered a ricepile model where the slope threshold $m_c$ is selected at random after each toppling. A modified version of this model was later used by Christensen {\em et al} \cite{christensen}. In a somehow different model Nunes-Amaral and Lauritsen \cite{nunes1} considered a rice pile model where the toppling rule is stochastic in certain range of slopes $m_{c1}<m<m_{c2}$ while it is deterministic above $m_{c2}$. All these ricepile models give the same avalanche exponents and should belong to the random disorder universality class. In table \ref{tab:1} we display the more accurate numerical estimates reported by Nunes-Amaral and Lauritsen. In $d=2$ we count with a critical slope model introduced by Bassler and Paczuski \cite{bassler} to describe the avalanche dynamics in superconducting vortexpiles. In their model $z_i$ is the density of vortex at site $i$. The evolution rules are not exactly those described above but a critical slope condition was used. They also considered quenched random pinning forces and therefore it should belong to the random disorder universality class. More recently Cruz, Mulet and Atshuler \cite{cruz} have made simulations of the model introduced by Bassler and Paczusky but considering both random and periodic pinning forces. In the case of random pinning their numerical estimates reproduced, within the numerical error, those reported by Bassler and Paczuski \cite{bassler}. However, in the case of periodic pinning they obtained very different exponents (see table \ref{tab:2}) suggesting that there are two universality class. Within our analysis the existence of these two universality classes is clear, corresponding with random and periodic distributions of pinning forces. In all cases we observe a very good agreement between the numerical estimates and our predictions. This agreement is more surprising because the dimensions considered are far from the upper critical dimension $d_c=4$. In table \ref{tab:1} we also display the scaling exponents $D$ and $z$ obtained from numerical simulations of the LID model driven at constant force, which are expected to be identical to those for the LID model driven at one end. As one can see the agreement is quite good, event better than with the RG estimates. Recently Tadi\'c and Nowak \cite{tadic} has observed that the scaling exponents of the random-field Ising model (RFIM) in the low disorder regime are very close to those of the ricepile model (see table \ref{tab:1}). In particular they considered a diluted two-dimensional Ising model with weak random fields and with an anisotropic initial condition. In the low disorder regime a single domain wall separating two regions with different spin orientations is observed. In this regime the Barkhausen avalanches are attributed to the fluctuations in the domain wall. The agreement with the exponents for the ricepile model suggests that the domain walls in this regime may be also described by the equation of motion \ref{eq:2}, while the anisotropy was introduced through the initial conditions. On the contrary, on the high disorder regime a multi-domain structure is obtained, resulting in different avalanche exponents. In this analysis we have not included a comparison with some available experimental results for sandpiles \cite{held}, ricepiles \cite{frette1} and vortexpiles \cite{superconductors} because it is very difficult to compare our results with experimental measurements. In most of the experiments the measured magnitude is the number of grains (vortexes) leaving (entering) the system which will be denoted by $s_0$. In general $s_0$ can be also described by the scaling law $P(s_0)\sim s_0^{-\tau_0}f(s_0/L^{D_0})$. However, in average, for one grain (vortex) entering the system one grain (vortex) goes out and therefore $\langle s_0\rangle=1$. Then if one computes $\langle s_0\rangle$ using the scaling law for $P(s_0)$ one obtains $\langle s_0\rangle\sim L^{(2-\tau_0)D_0}\sim1$, which immediately gives $\tau_0=2$. This value is near the one reported for real sandpiles, ricepiles and vortex piles. On the contrary, in numerical simulations and in the model analyzed in this work the avalanche size $s$ is given by the number of toppling events which is not necessarily proportional to the flow of grains (vortexes), the magnitude measured in experiments. In fact, in the case of the avalanche size $s$, the conservation law yields $\tau_s=2-1/D<2$) (see eq. (\ref{eq:15}). Hence, we conclude that $\tau_0\neq\tau_s$. \section{conclusions} We conclude that ricepile models, the vortexpile model by Bassler and Paczuski, and the RFIM in the low disorder regime belong to the same universality class, that of critical slope sandpile models. The prototype model for this universality class is the LID model dragged at one end, as it was already conjectured by Paczuski and Boettcher \cite{paczuski}. Our analytical treatment has demonstrated that their guess was correct. For the first time, we have solved the LID model dragged at one end, obtaining the complete set of scaling exponents. Some of these results were supported by previous RG calculations developed for constant force LID but others, like the linear dependence of the susceptibility with system size, are intrinsic to this model. Moreover, as in the constant force case, we found two different universality classes corresponding with random and periodic pinning forces. Our predictions were found in very good agreement with numerical simulations of different critical slope models. \section*{ackowledgements} This work has been partly supported by the {\em Alma Mater } prize, from the Havana University.
\section{Introduction} Scalar-tensor (ST) theories of gravity have an extensive history \cite{jr,fr,bd,brg,ndt,wag1,dm-ef}. Their study is important for a number of reasons, in particular because they provide the simplest generalization of Einstein's Theory of General Relativity (GR); they also turn out to be the low energy limit of certain attempts at a quantum theory of gravity, such as superstrings \cite{ss}. In the present paper we review ST gravity in the physical or Jordan frame and the conformally related Einstein frame. The formulation of ST theory in the physical frame (sec. II) apparently does not lead to a well defined energy-momentum tensor for the scalar field. The scalar field terms on the right of the Einstein field equation, the quantities that one would naively associate with the energy-momentum tensor of the scalar field, prove to be ill behaved to constitute such a tensor: in the first place, the scalar field energy density cannot be made universally nonnegative, as the presence of second derivatives makes it impossible. This has led some authors (see \cite{far,mag} and references therein) to reject the physical frame on exactly the grounds of the undesirable features of this ``apparent'' energy-momentum tensor. However, a careful look at it allows us to conclude that {\it all} scalar field terms on the right hand side of the Einstein equation {\it should not be identified with the energy-momentum tensor of the scalar field}. In fact, the terms with the second covariant derivatives of the scalar field contain the connection, and hence a part of the dynamical description of gravity. This assertion is substantiated by the origin of the second derivative terms: they come from variation of the gravitational part of the action with respect to the metric. We find a new connection that describes the correct dynamics of gravity in sec III. The description in terms of this new connection removes the gravitational dynamical terms from the right of the Einstein equation, leaving us with the correct energy-momentum tensor for the scalar field, the one that is not polluted by gravitational dynamical terms. The scalar field energy density can now be made nonnegative, which condition is implemented in the form of two important inequalities for the two otherwise arbitrary functions that define a particular ST theory (the coupling function and the potential, i.e., the scalar field dependent cosmological "constant"). Moreover, the new connection arises from a metric conformally related to the physical metric. The conformal frame of this new metric is the Einstein frame. For completeness we develop ST gravity in the Einstein frame in sections IV and V of the paper. Our main motivation for obtaining the described results was a serious confusion in and long discussion of the meaning and role of the Jordan and Einstein frames. For a good review of the different points of view and a very thorough list of references we refer readers to the article by Faraoni, \emph{et. al.} \cite{far}. The confusion is about what frame should be considered the "true physical" one. The source of the confusion is the fact that the Einstein frame metric is the one that correctly describes the spin-2 dynamics of gravity. For this reason some authors single out the Einstein frame as the only one fundamental. On the other hand, by construction the metric in the Jordan frame is the one that determines metrical relations in spacetime, and particles move on geodesics of the physical metric, so the Jordan frame is claimed to be the fundamental one on these grounds. We think that no such recognition is pertinent, since it depends on whether one ascribes more importance to the dynamics or to metrical relations and geodesics. There is even a belief that the two frames lead to different physics (see \cite{far,mag} and references therein). In our opinion, this statement should be taken only to the extent that the physical and Einstein frame metrics behave differently. Apart from that, those are just two different descriptions of the same physics. Physical properties of the continuum are determined by the physical frame metric because this is the metric to which nongravitational fields couple universally by construction. Otherwise, since there is a well defined transformation between the two conformally related frames, one can work in whatever frame is convenient as long as one uses the physical metric in the end to describe the physics of a particular problem. The authors follow the basic conventions of Misner, Thorne, and Wheeler \cite{mtw} throughout the paper. \label{sec-int} \section{Scalar-Tensor Theories of Gravity in the Physical Frame} We consider the most general scalar-tensor (ST) theories of gravity with a single scalar field. In these theories the gravitational interaction is mediated by the metric $\tilde{g}_{\mu \nu}$ and a spin-0 field, a scalar field $\Phi$. The field equations for these theories follow from the action \cite{jr,fr,bd,brg,ndt,wag1,dm-ef,dm-nd} \begin{eqnarray} S=\frac{1}{16\pi}\int d^{4}x\sqrt{-\tilde{g}}\,[\Phi \tilde{R} - \frac{\omega(\Phi)}{\Phi}\tilde{g}^{\mu \nu}\Phi_{, \mu}\Phi_{, \nu} -2\tilde{\Lambda}(\F)] \nonumber \\ +\;\; S_{m}[\Psi_{m},\tilde{g}_{\mu \nu}] \label{eq:aph} , \end{eqnarray} where $ \ _{,\mu}$ represents the partial derivative with respect to $x^{\mu}$, $\tilde{R}$ is the Ricci scalar constructed from the metric $\tilde{g}_{\mu \nu}$, $\omega(\Phi)$ is the coupling function of the scalar field to matter, and the cosmological term $\tilde{\Lambda}(\F)$ is the scalar field potential. The scalar field $\Phi$ plays the role of the inverse gravitational constant $G^{-1}$. To ensure that gravity be attractive we impose the condition \begin{equation} \Phi > 0 \end{equation} The last term in (\ref{eq:aph}) is the action of the matter fields, $\Psi_{m} $, which couple only to the metric $\tilde{g}_{\mu \nu}$ and not to the scalar field $\Phi$, in order to satisfy the weak equivalence principle. This formulation of ST gravity is called the physical (or Jordan) frame description~\cite{dm-ef,dm-nd}, because the metric in this frame is the ``true'' metric of our spacetime. By ``true'' we mean that this metric is the one measured by standard rods and clocks, i. e., it is the one that determines the geometry of our spacetime. The proper time measured by a moving test particle is given by $d\tilde{\tau}^2 = \tilde{g}_{\mu \nu} dx^\mu dx^\nu$. The 4-velocity, $\tilde{u}^\mu= dx^\mu /d\tilde{\tau}$, of the particle satisfies the geodesic equation \begin{equation} \tilde{u}^{\mu}\, _{, \beta}\tilde{u}^\beta +\tilde{\Gamma}^{\mu}_{\nu \beta} \tilde{u}^\nu \tilde{u}^\beta=0 \label{eq:geo} \, , \end{equation} where the connection $\tilde{\Gamma}^{\mu}_{\nu \beta}$ is the Christoffel symbol calculated with respect to the physical metric $\tilde{g}_{\mu \nu}$. So far we only assume that $\omega(\Phi)$ and $\tilde{\Lambda}(\F)$ are smooth enough functions on the positive semiaxis. The fact that $\omega(\Phi)$ could be negative might seem puzzling since the action (\ref{eq:aph}) would then appear to imply a negative kinetic term for the scalar field energy density. This is not the case because in the physical frame it is not possible to define a suitable energy density for the scalar field due to its nonminimal coupling (through $\Phi \tilde{R}$) to the gravitational part of the action. Hence there is no suitable geometrical definition of the energy-momentum tensor of the scalar field. The spin-2 and spin-0 excitations are ``entangled'' in the physical frame. In the following section we proceed to ``disentangle'' the two propagation modes by suitable transformations. Then it becomes possible to define the energy-momentum tensor and the energy density for the scalar field; the non-negativity of the latter requires $\omega(\Phi)$ and $\tilde{\Lambda}(\F)$ to satisfy some inequalities. Each particular choice of the two arbitrary functions $\omega(\Phi)$ and $\tilde{\Lambda}(\F)$ specifies a different ST theory of gravity. In general, ST gravity represents theories with cosmological and gravitational ``constants'' which change from point to point in spacetime. We vary the action (\ref{eq:aph}) with respect to $\tilde{g}_{\mu \nu}$ and $\Phi$ to obtain the field equations \begin{eqnarray} \tilde{G}_{\mu \nu}=8\pi \frac{\tilde{T}_{\mu \nu}}{\Phi} + \frac{\omega(\Phi)}{\Phi^2}( \Phi_{,\mu}\Phi_{, \nu} - 1/2 \tilde{g}_{\mu \nu} \tilde{g}^{\alpha \beta} \Phi_{,\alpha} \Phi_{,\beta}) -\tilde{g}_{\mu \nu} \frac{\tilde{\Lambda}(\F)}{\Phi} \nonumber \\ +\frac{1}{\Phi}(\tilde{\nabla_{\mu}} \tilde{\nabla_{\nu}} \Phi - \tilde{g}_{\mu \nu} \Box_{\tilde{ g}} \Phi) \label{eq:tG}\\ \Box_{\tilde{g}}\Phi + \frac{1}{2}\Phi_{, \alpha} \Phi_{, \beta} \tilde{g}^{\alpha \beta} \frac{d}{d\Phi} \ln \left(\frac{\omega(\Phi)}{\Phi} \right) + \frac{\Phi}{2 \omega (\Phi)} \left[ \tilde{R} - 2 \frac{d \tilde{\tilde{\Lambda}(\F)}}{d \Phi}\right] =0 \label{eq:F1}\, \end{eqnarray} Here $\tilde{\nabla}_{\mu}$ is the covariant derivative with respect to $\tilde{g}_{\mu \nu}$, $\Box_{\tilde{g}}\Phi= [\sqrt{-\tilde{g}}\tilde{g}^{\mu \nu} \Phi_{, \mu}]_{,\nu}/ \sqrt{-\tilde{g}}$ is the covariant D'Alambertian, $\tilde{G}_{\mu \nu} =\tilde{R}_{\mu\nu}-\left(1/2 \right)\tilde{g}_{\mu \nu} \tilde{R}$ is the Einstein tensor, and $\tilde{R}_{\mu\nu}$ is the Ricci tensor. By contracting equation (\ref{eq:tG}) and using the result to remove $\tilde{R}$ from the scalar field equation (\ref{eq:F1}) we find \begin{equation} \left[2 \omega (\Phi) + 3\right]\Box_{\tilde{g}}\Phi = 8 \pi \tilde{T} + 2 \Phi \frac{d\tilde{\Lambda}(\F)}{d \Phi} - 4 \tilde{\Lambda}(\F) - \frac{d \omega (\Phi)}{d \Phi} \Phi_{, \alpha} \Phi_{, \beta} \tilde{g}^{\alpha \beta}, \label{eq:F} \end{equation} where $\tilde{T}_{\mu \nu}$ is the physical frame energy-momentum tensor of matter defined by the usual geometrical expression via the variational derivative as \begin{equation} \tilde{T}_{\mu \nu}=-\frac{2}{\sqrt{-\tilde{g}}}\frac{\delta S_{m}}{\delta \tilde{g}^{\mu \nu}} \, \label{eq:tT} \end{equation} Note that the values of the scalar field at which the coupling function $\omega (\Phi)$ turns to $-3/2$ are singular points of the scalar field equation (\ref{eq:F}). Since $\tilde{g}_{\mu \nu}$ is the metric which matter ``feels'', $\tilde{T}_{\mu \nu}$ is the ``true'' energy-momentum tensor. Again, by ``true'' we mean that physically measurable quantities are the ones related to it. For example, an observer with the 4-velocity $\tilde{u}^\mu$ would measure the energy density $\epsilon = \tilde{T}_{\mu \nu} \tilde{u}^\mu \tilde{u}^\nu$. Since the weak equivalence principle is satisfied (the matter fields only couple to $\tilde{g}_{\mu \nu}$), the energy-momentum tensor of the matter fields is conserved: \begin{equation} \tilde{\nabla}_{\nu} \tilde{T}^{\mu \nu}=0. \label{eq:pcons} \end{equation} The direct derivation of this conservation law from equations (\ref{eq:tG}) and (\ref{eq:F1}) is not at all straightforward. \section{Energy-Momentum Tensor of the Scalar Field} Let us now take a careful look at the Einstein field equation (\ref{eq:tG}). One usually associates the quantities on the right hand side with the energy-momentum tensor of the matter and non-gravitational physical fields. As mentioned above, $\tilde{T}_{\mu \nu}$ is the energy-momentum tensor of the matter fields as it comes from the variation of the matter action with respect to $\tilde{g}_{\mu \nu}$. One is tempted to identify the scalar field terms with the energy-momentum tensor of the scalar field, but the situation is not that simple. On the right of equation (\ref{eq:tG}) there are terms which depend on the scalar field $\Phi$ and its first derivatives, and terms which are linear in the second order covariant derivatives of $\Phi$. The former terms, \begin{equation} \frac{\omega(\Phi)}{\Phi^2}( \Phi_{,\mu}\Phi_{,\nu} - 1/2 \tilde{g}_{\mu \nu} \tilde{g}^{\alpha \beta} \Phi_{,\alpha} \Phi_{, \beta}) -\tilde{g}_{\mu \nu} \frac{\tilde{\Lambda}(\F)}{\Phi} \, , \label{eq:3.1} \end{equation} come from varying the purely scalar field parts of the action with respect to $\tilde{g}_{\mu \nu}$, that is, from varying everything that contains $\Phi$, except $\Phi \tilde{R}$ which describes some of the dynamics of the gravitational field. Hence expression (\ref{eq:3.1}) can be identified with a part of the scalar field energy-momentum tensor of the scalar field. The remaining scalar field terms on the right of equation (\ref{eq:tG}), those that have second covariant derivatives in them, $$ \frac{1}{\Phi}(\tilde{\nabla_{\mu}} \tilde{\nabla_{\nu}} \Phi - \tilde{g}_{\mu \nu} \Box_{\tilde{ g}} \Phi)\, , $$ should not belong to the scalar field energy-momentum tensor. First of all, the presence of second derivatives of $\Phi$ is undesirable because it would make it impossible to have a nonnegative energy density for the scalar field. Even more important, these terms contain a combination $$ \frac{1}{\Phi}(\tilde{g}_{\mu \nu} \tilde{\Gamma}^{\alpha}_{\sigma \alpha}\tilde{g}^{\sigma \beta }\Phi_{\beta} - \tilde{\Gamma}^{\alpha}_{\mu \nu}\Phi_{\alpha}) \, $$ with the connection in it. It comes, after integration by parts, from varying $\Phi \tilde{R}$ which carries the dynamical information on the tensor part of gravity. Hence these terms should be regarded as a part of the dynamical evolution and constraint equations for the gravitational field. Therefore these terms may belong on the left hand side of equation (\ref{eq:tG}) where the rest of the dynamical description of the gravitational field (in the form of the Einstein tensor $\tilde{G}_{\mu \nu}$) resides. In order to have all the terms that contain the dynamical description of gravity on the left hand side of the Einstein equation (\ref{eq:tG}) we define a new connection \begin{equation} \Gamma^{\alpha}_{\beta \gamma} = \tilde{\Gamma}^{\alpha}_{\beta \gamma} + \frac{1}{2\Phi} \left( \delta^{\alpha}_{\beta} \Phi_{, \gamma} + \delta^{\alpha}_{ \gamma} \Phi_{, \beta} - \tilde{g}_{\beta \gamma} \Phi^{, \alpha} \right) \, , \label{eq:con} \end{equation} ($\Phi^{,\alpha} \equiv \tilde{g}^{\alpha \sigma} \Phi_{, \sigma}$). That $\Gamma^{\alpha}_{\beta \gamma}$ transforms as a connection follows from the fact that \begin{equation} D^{\alpha}_{\beta \gamma} \equiv \frac{1}{2 \Phi}\left( \delta^{\alpha}_{\beta} \Phi_{, \gamma} +\delta^{\alpha}_{\gamma} \Phi_{, \beta} - \tilde{g}_{\beta \gamma} \Phi^{, \alpha} \right) \label{eq:D} \end{equation} transforms as a $(1,2)$ tensor. Since $\tilde{\Gamma}^{\alpha}_{\beta \gamma}$ is the connection compatible with the metric ($\tilde{\nabla}_{\alpha} \tilde{g}_{\mu \nu} =0$), the connection $\Gamma^{\alpha}_{ \beta \gamma}$ is generally not compatible with the metric. Indeed, one easily finds $\nabla_{\alpha} \tilde{g}_{\mu \nu} = - \tilde{g}_{\mu \nu} \Phi_{,\alpha}/ \Phi$, where $\nabla$ represents the covariant derivative with respect to the connection $\Gamma^{\alpha}_{\beta \gamma}$. In the Appendix we derive the following relation $(R_{\mu \nu}$ is the Ricci tensor calculated from the connection $\Gamma^{\alpha}_{\beta \gamma} $ ): \begin{equation} \tilde{R}_{\mu \nu} = R_{\mu \nu} + \frac{1}{\Phi} \left( \tilde{\nabla_{\mu}} \tilde{\nabla_{\nu}} \Phi + \frac{1}{2}\tilde{g}_{\mu \nu} \Box_{\tilde{g}} \Phi \right) - \frac{3}{2} \frac{\Phi_{,\mu} \Phi_{,\nu}}{\Phi^2} \, \label{eq:3.4} \end{equation} If we now define the Ricci scalar and Einstein tensor with respect to the new connection as \begin{equation} R^* \equiv \tilde{g}^{\mu \nu} R_{\mu \nu}, \qquad G_{\mu \nu} \equiv R_{\mu \nu} - \frac{1}{2} \tilde{g}_{\mu \nu} R^* \, , \end{equation} we then obtain: \begin{eqnarray} \tilde{R} = R^* + \frac{3 \Box_{\tilde{g}} \Phi}{\Phi} - \frac{3}{2} \frac{\Phi^{, \alpha} \Phi_{,\alpha}}{\Phi^2}\, , \label{eq:ricc}\\ \tilde{G}_{\mu \nu} = G_{\mu \nu} + \frac{1}{\Phi}(\tilde{\nabla_{\mu}} \tilde{\nabla_{\nu}} \Phi - \tilde{g}_{\mu \nu} \Box_{\tilde{g}} \Phi) \nonumber \\ - \frac{3}{2 \Phi^2} ( \Phi_{,\mu}\Phi_{,\nu} - 1/2 \tilde{g}_{\mu \nu} \tilde{g}^{\alpha \beta} \Phi_{,\alpha} \Phi_{,\beta} ) \label {eq:Eins}\, . \end{eqnarray} One immediately sees that the right hand side of the last equation includes the ``offending'' second covariant derivative terms, which will cancel when we substitute (\ref{eq:Eins}) in the Einstein equation (\ref{eq:tG}). In terms of the new Einstein tensor, $G_{\mu \nu}$, the latter becomes \begin{equation} G_{\mu \nu}=8\pi \frac{\tilde{T}_{\mu \nu}}{\Phi} + \frac{\omega(\Phi) + 3/2}{\Phi^2}( \Phi_{,\mu} \Phi_{,\nu} - 1/2 \tilde{g}_{\mu \nu} \tilde{g}^{\alpha \beta} \Phi_{,\alpha} \Phi_{,\beta}) -\tilde{g}_{\mu \nu} \frac{\tilde{\Lambda}(\F)}{\Phi} \label{eq:G0} \, . \end{equation} We see that we have succeeded in eliminating the terms containing the connection on the right hand side with an additional bonus that the full second covariant derivatives of the scalar field were eliminated. The dynamical description of gravity is in terms of the connection $\Gamma^{\alpha}_{\beta \gamma}$ and not the metric connection $\tilde{\Gamma}^{\alpha}_{\beta \gamma} $. Hence, $G_{\mu \nu}$ describes the complete dynamics of the gravitational field and, therefore, we can identify the quantities to its right in equation (\ref{eq:G0}) as the energy-momentum tensor of the matter and physical fields. We have effectively ``disentangled'' the tensor and scalar modes. In particular the energy-momentum tensor, $\tilde{\Sigma}_{\mu \nu}$, of the scalar field is given by \begin{equation} 8 \pi \frac{\tilde{\Sigma} _{\mu \nu}}{\Phi} \equiv \frac{\omega(\Phi) + 3/2}{\Phi^2} ( \Phi_{,\mu}\Phi_{,\nu} - 1/2 \tilde{g}_{\mu \nu} \tilde{g}^{\alpha \beta} \Phi_{,\alpha} \Phi_{,\beta}) -\tilde{g}_{\mu \nu} \frac{\tilde{\Lambda}(\F)}{\Phi} \,; \label{eq:Sig} \end{equation} note that unlike the energy-momentum tensor of matter, $\tilde{T}_{\mu \nu}$ , it is not, of course, covariantly conserved: $$ \tilde{\nabla}_{\mu} \tilde{\Sigma}^{\mu}_{\nu} = \frac{\Phi_{,\nu}}{2\Phi} \left[ \tilde{T} - \frac{\tilde{\Lambda}(\F)}{2 \pi} - \frac{\omega(\Phi) + 3/2}{8 \pi \Phi}\Phi_{,\alpha} \Phi^{,\alpha}\right] \not\equiv0 $$ This nonconservation is a consequence of the nonminimal coupling between the scalar field and the metric. We now impose the condition that the energy density $\tilde{\Sigma} _{\mu \nu} \tilde{u}^\mu \tilde{u}^\nu$ of the scalar field for any observer be nonnegative. Expression (\ref{eq:Sig}) shows that this can be achieved if and only if \begin{eqnarray} \omega(\Phi) \ge -\frac{3}{2} \label{eq:om}\\ \tilde{\Lambda}(\F) \ge 0 \label{eq:lam} \, ; \end{eqnarray} we consider these inequalities valid for our ST gravity. The classical treatment of the field equations requires the functions $\omega(\Phi)$ and $\tilde{\Lambda}(\F)$ to have at least one continuous derivative for $\Phi>0$, which condition we always assume in this paper. Note that if we define a new metric, $g_{\mu \nu}$, by the conformal transformation $g_{\mu \nu} = \Phi \tilde{g}_{\mu \nu}$, then the Christoffel symbol $\{ ^{\; \alpha}_{\beta \gamma} \}$ with respect to $g_{\mu \nu}$, i. e., the connection compatible with $g_{\mu \nu}$, satisfies \begin{equation} \left\{ ^{\; \alpha}_{\beta \gamma} \right\} = \tilde{\Gamma}^{\alpha}_{\beta \gamma} + \frac{1}{2\Phi} \left( \delta^{\alpha}_{\beta} \Phi_{, \gamma} + \delta^{\alpha}_{\gamma} \Phi_{, \beta} - \tilde{g}_{\beta \gamma} \Phi^{, \alpha} \right) \, . \label{eq:3.12} \end{equation} By (\ref{eq:3.12}) and (\ref{eq:con}) we have $\Gamma^{\alpha}_{\beta \gamma} = \{ ^{\; \alpha}_{\beta \gamma} \}$; therefore $\Gamma^{\alpha}_{\beta \gamma}$ is the metric connection for $g_{\mu \nu}$. We conclude that in ST gravity there is a metric $\tilde{g}_{\mu \nu}$ which determines proper lengths and times, and geodesics in our spacetime, i. e., it is a metric in the proper sense of the word. There is also the conformally related ``metric'' $g_{\mu \nu}$ which carries the dynamical information of the gravitational field, in other words it describes the pure spin-2 excitations. However, it is important to bear in mind that the expression $g_{\mu \nu} dx^\mu dx^\nu$ does not represent a physical spacetime interval. We call $g_{\mu \nu}$ the dynamical metric and $\Gamma^{\alpha}_{\beta \gamma}$ the dynamical connection. The corresponding frame conformally related to the physical frame is called the Einstein frame \cite{dm-ef,dm-nd,far}. \section{Scalar-Tensor Theories of Gravity in the Einstein Conformal Frame} We now proceed to develop the description of ST gravity in the Einstein frame. As we mentioned, this is the description in terms of the conformally related dynamical metric and connection. Instead of working from the field equations, it is more convenient to work from the action (\ref{eq:aph}). The only dynamical geometrical quantity in the action is the Ricci scalar $\tilde{R}$. In the previous section we derived an expression for this quantity in terms of the Ricci scalar $R^*$ defined by contracting the dynamical Ricci scalar $R_{\mu \nu}$ with the physical metric $\tilde{g}^{\mu \nu}$. The appropriate Ricci scalar for the Einstein frame description is defined by contracting with the dynamical metric: \begin{equation} R= g^{\mu \nu} R_{\mu \nu}= \frac{\tilde{g}^{\mu \nu}}{\Phi} R_{\mu \nu}= \frac{R^*}{\Phi} \, . \end{equation} In terms of $R$, equation (\ref{eq:ricc}) becomes \begin{equation} \tilde{R} = \Phi R + \frac{3 \Box_{\tilde{g}} \Phi}{\Phi} - \frac{3}{2} \tilde{g}^{ \alpha \beta} \frac{\Phi_{,\alpha} \Phi_{,\beta}}{\Phi^2}\, . \end{equation} The quantities on the right hand side have to be expressed in terms of the dynamical metric. Since $g_{\mu \nu} = \Phi \tilde{g}_{\mu \nu}$ we have $$ \tilde{g}^{\mu \nu} = \Phi g^{\mu \nu},\qquad \sqrt{-\tilde{g}}=\frac{\sqrt{-g}}{\Phi^2}. \qquad \Box_{\tilde{g}} \Phi =\frac{1}{\sqrt{-\tilde{g}}} \left[ \sqrt{-\tilde{g}} \tilde{g}^{\mu \nu} \Phi_{,\mu}\right]_{,\nu} = \frac{\Phi^2}{\sqrt{-g}} \left[ \frac{\sqrt{-g}}{\Phi} g^{\mu \nu} \Phi_{,\mu}\right]_{,\nu} = \Phi^2 \Box \ln \Phi \, . $$ Using these relations we obtain \begin{equation} \tilde{R} = \Phi R + 3 \Phi \Box \ln \Phi - \frac{3}{2} \frac{\Phi^{,\alpha} \Phi_{, \alpha}}{\Phi}\, , \label{eq:ricci} \end{equation} where from now on all indices are raised and lowered with the dynamical metric $g_{\mu \nu}$, i. e. $\Phi^{, \mu} = g^{\mu \nu} \Phi_{,\nu}$, unless otherwise indicated. The action (\ref{eq:aph}) in terms of the dynamical metric becomes \begin{eqnarray} S=\frac{1}{16\pi}\int d^{4}x\sqrt{-g}\,[ R - \frac{\omega(\Phi) + 3 /2 }{\Phi^2} g^{\mu \nu}\Phi_{, \mu}\Phi_{, \nu} -2\frac{\tilde{\Lambda}(\F)}{\Phi^2}] + S_{m}[\Psi_{m},g_{\mu \nu} / \Phi] \label{eq:a} \, , \end{eqnarray} where the term containing $3 \Box \ln \Phi$ was integrated by parts to give zero by supposing all our fields vanish on the boundary of spacetime. Let us pause briefly to mention a few properties of our Einstein frame action (\ref{eq:a}). One immediately notices that the metric and scalar field parts are now untangled in the sense that no scalar field dependent factor stands in front of $R$ in (\ref{eq:a}). The dynamics of gravity is described by the Ricci scalar $R$, which now appears by itself without any scalar field dependent factors. There is also the scalar field part of the action in which the only coupling to gravity is through the metric; it leads thus to a well defined energy-momentum tensor for the scalar field. Moreover, it is clear that the two conditions (\ref{eq:om}) and (\ref{eq:lam}) must be satisfied, to have a nonnegative scalar field energy density. On the other hand, the separation of the dynamics of the metric and scalar field comes with a price: matter couples to gravity nonminimally (but universally) through the physical metric $\tilde{g}_{\mu \nu} = g_{\mu \nu} / \Phi$, to preserve the weak equivalence principle. The Einstein frame field equations follow by varying the action (\ref{eq:a}). When it is varied with respect to the dynamical metric $g^{\mu \nu}$, the Einstein field equation \begin{equation} G_{\mu \nu}=8\pi T_{\mu \nu} + \frac{\omega(\Phi) + 3/2}{\Phi^2}( \Phi_{,\mu} \Phi_{,\nu} - 1/2 g_{\mu \nu} g^{\alpha \beta} \Phi_{,\alpha} \Phi_{,\beta}) -\tilde{g}_{\mu \nu} \frac{\tilde{\Lambda}(\F)}{\Phi^2} \equiv 8\pi\left( T_{\mu \nu}+\Sigma_{\mu \nu}\right) \label{eq:G} \end{equation} is obtained. The Einstein tensor on its left describes the evolution and constraints for the dynamical metric. In the right hand side are the sources, namely, the energy-momentum tensor of the scalar field defined in (\ref{eq:Sig}) and the Einstein frame energy-momentum tensor, $T_{\mu \nu}$, of the matter fields whose standard definition is \begin{equation} T_{\mu \nu}=-\frac{2}{\sqrt{-g}}\frac{\delta S_{m}}{\delta g^{\mu \nu}} \, . \label{eq:T} \end{equation} We stress that the Einstein frame energy-momentum tensor does not represent the energy-momentum tensor of the matter fields, because it is defined by the variation with respect to the dynamical metric which is not the metric of the physical continuum. As we mentioned earlier, the physical frame energy-momentum tensor is the ``true'' one as it is the physical metric that defines metrical relations in spacetime. From the definitions of the energy-momentum tensors of matter in each frame ( see equations (\ref{eq:tT}) and (\ref{eq:T})) the relations between their components follow easily: \begin{equation} T_{\mu \nu} = \frac{\tilde{T}_{\mu \nu}}{\Phi},\qquad T^{\mu}_{\nu} = \frac{\tilde{T}^{\mu}_{\nu}}{\Phi^2},\qquad T^{\mu \nu} = \frac{\tilde{T}^{\mu \nu}}{\Phi^3},\qquad T \equiv T^{\mu}_{\mu} = \frac{\tilde{T}^{\mu}_{\mu}}{\Phi^2} \equiv \frac{\tilde{T}}{\Phi^2} \end{equation} Here indices for the Einstein frame tensor are raised and lowered with the dynamical metric $g_{\mu \nu}$ and for the physical frame tensor with the physical metric $\tilde{g}_{\mu \nu}$. The conservation of the physical energy-momentum tensor (\ref{eq:pcons}) leads to the following differential equation for the Einstein frame tensor: \begin{equation} T^{\mu}_{\nu}\,_{; \mu}= - \frac{\Phi_{, \nu}}{2\Phi} T \, , \label{eq:cons} \end{equation} where $;$ represents covariant differentiation with respect to the dynamical metric. The Einstein frame energy-momentum tensor is not covariantly conserved, hence free particles do not follow geodesics of the dynamical metric. The wave equation for the scalar field is obtained by varying the action (\ref{eq:a}) with respect to $\Phi$: \begin{equation} \left[\omega(\Phi) +3/2\right]\left(\Box \Phi - \frac{\Phi_{, \alpha} \Phi^{, \alpha}}{ \Phi}\right) ={\Phi^2}\left\{ \frac{4 \pi T}{\Phi} + \frac{d}{d \Phi} \left[ \frac{ \tilde{\Lambda}(\F)}{\Phi^2} \right]\right\}-\frac{1}{2}\Phi_{, \alpha} \Phi^{, \alpha}\frac{d \omega(\Phi)}{d\Phi} \label{eq:Phi2} \end{equation} Using this we demonstrate that the scalar field energy-momentum tensor is not covariantly conserved: \begin{equation} \Sigma^{\mu}_{\nu}\,_{; \mu} = \frac{\Phi_{, \nu}}{2\Phi} T \, . \label{eq:Phicons} \label{eq:S} \end{equation} However, from (\ref{eq:cons}) and (\ref{eq:S}) it is clear that the sum of the matter and scalar field energy-momentum tensors is covariantly conserved as it must because of the contracted Bianchi identity, $G^{\mu}_{\nu}\,_{; \mu}=0$. Note that in the Einstein frame we have an appropriate description of the tensor part of gravity and of the scalar field. Moreover the energy-momentum tensor for the scalar field can only be defined in terms of the dynamical connection and/or dynamical metric. On the other hand, it costs a more complicated description of matter, with a nonconserved energy-momentum tensor. In the physical frame the description of gravity and of the scalar field are complicated since their propagation modes are entangled, but the description of matter is simple. Each frame has its technical advantages, but one must remember that it is the physical metric and energy-momentum tensor which are most directly related to observables. \section{ Scalar Field Redefinition in the Einstein Frame} As stated in the previous section, the Einstein frame provides the description of our gravitational theory in terms of the dynamical metric $g_{\mu \nu}= \Phi \tilde{g}_{\mu \nu}$. One has yet a freedom to redefine the scalar field, which choice is usually used to simplify the scalar field equation (\ref{eq:Phi2}), complicated so far by the presence of terms quadratic in the scalar field derivatives. To do that, one realizes that the latter originate from the term \begin{equation} \frac{\omega(\Phi) + 3 /2 }{\Phi^2} g^{\mu \nu}\Phi_{, \mu}\Phi_{, \nu} \, . \label{eq:5.1} \end{equation} in the action (\ref{eq:a}). If the factor in front of the derivative terms were a constant, the scalar field equation would simplify significantly. In view of the inequality (\ref{eq:om}) imposed on the coupling function $\omega(\Phi)$, this immediately gives rise to the two following cases. \medskip Case 1. $\omega(\Phi)\equiv -3/2$ The coefficient in (\ref{eq:5.1}) is not just a constant, but exactly zero, so no field redefinition is needed. This is a rather peculiar situation from the physical standpoint, since both the energy momentum tensor of the scalar field (\ref{eq:Sig}) and the Einstein frame action (\ref{eq:a}) contain no kinetic terms at all. As a consequence the scalar field equation (\ref{eq:Phi2}) reduces to \begin{equation} \frac{4 \pi T}{\Phi} + \frac{d}{d \Phi} \left[ \frac{\tilde{\Lambda}(\F)}{\Phi^2} \right] =0 \label{eq:5.2} \end{equation} Instead of being an evolution equation, (\ref{eq:5.2}) is rather an algebraic equation for the scalar field $\Phi$ which allows for its direct determination from the matter term by means of inversion. The physical plausibility of this case seems questionable to us. \medskip Case 2. $\omega(\Phi)\not\equiv -3/2$ In this case we define a new scalar field $\chi=\chi(\Phi)$ by the equations ($\Phi_0>0$ is a constant) \begin{equation} \sqrt{2} \frac{d \chi}{d \ln \Phi} = \sqrt{\omega(\Phi) + 3 /2},\qquad \chi(\Phi)=\frac{1}{\sqrt2}\int_{\Phi_0}^\Phi \frac{\sqrt{\omega(\Phi) + 3 /2}}{\Phi}\,d\Phi, \label{eq:defchi} \end{equation} so that the coefficient in (\ref{eq:5.1}) turns into $2$ after replacing $\Phi$ by $\chi$. It is clear that the new scalar field $\chi$ is a non-decreasing function of the old one, $\Phi\geq0$, and its growth can stop only at such values $\Phi_*$ for which $\omega(\Phi_*)=-3/2$; if they exist, those are evidently the inflection points of $\chi(\Phi)$. Therefore an inverse to $\chi(\Phi)$ function $\Phi=\Phi(\chi)$ is uniquely determined; it is a monotonically increasing function of its argument. The range of the new scalar field $\chi(\Phi)$ depends on the value of the coupling function at $\Phi=+0$. If $\omega(+0)= -3/2$ and the integral in (\ref{eq:defchi}) converges at $\Phi=0$, then the range is $$ -\infty<\chi(+0)<0\leq\chi(\Phi)\leq\chi(+\infty)\leq+\infty,\qquad 0\leq\Phi\leq+\infty, $$ If, on the other hand, $-3/2<\omega(+0)\leq\infty$, then the integral in (\ref{eq:defchi}) diverges at $\Phi=0$, and the range includes the negative semiaxis: $$ -\infty=\chi(+0)\leq\chi(\Phi)\leq\chi(+\infty)\leq+\infty, \qquad 0\leq\Phi\leq+\infty, $$ The upper limit $\chi(+\infty)$ of the range is either infinite or finite positive depending on whether the integral in (\ref{eq:defchi}) diverges or converges at $\Phi=+\infty$. For the second possibilty to occur, the coupling function must tend to $-3/2$ at infinity, so that the generic range of $\chi(\Phi)$, without any additional assumptions about $\omega(\Phi)$, is the whole real axis. \smallskip After this redefinition of the scalar field the Einstein frame action and field equations become \begin{equation} S=\frac{1}{16\pi}\int d^{4}x\sqrt{-g}\,[ R - 2 g^{\alpha \beta} \chi_{, \alpha} \chi_{, \beta} -2\Lambda(\chi)] + S_{m}[\Psi_{m},g_{\mu \nu} /\Phi (\chi)] \label{eq:as} \end{equation} \begin{equation} G_{\mu \nu}=8\pi T_{\mu \nu} + 2( \chi_{, \mu} \chi_{, \nu} - 1/2 g_{\mu \nu} g^{\alpha \beta} \chi_{, \alpha} \chi_{, \beta}) -g_{\mu \nu} \Lambda(\chi) \label{eq:Gs} \end{equation} \begin{equation} \Box \chi = \frac{1}{2} \frac{d\Lambda(\chi)}{d \chi} + \left[ 2 \pi \frac{d \ln \Phi(\chi)}{d \chi}\right]T , \label{eq:chi} \end{equation} where the function $\Lambda(\chi)$ is defined as \begin{equation} \Lambda(\chi) = \tilde{\Lambda}(\Phi (\chi))/\Phi^2(\chi), \label{eq:5.9} \end{equation} and $T_{\mu \nu}$ is the Einstein frame energy-momentum tensor of matter (\ref{eq:T}) (its interpretation and relations to the physical energy-momentum tensor were discussed in the previous section). The scalar field equation (\ref{eq:chi}) has the same singular points as the equation (\ref{eq:F}), namely, the values of $\Phi$ at which the coupling function turns to $-3/2$, because $d \Phi(\chi)/d \chi=\infty$ at these points. Note that in principle one can, of course, use the negative branch of the square root in the definition (\ref{eq:defchi}) of the function $\chi(\Phi)$, or even combine the positive and negative branches (to keep the derivative $d \chi /d \Phi$ continuous, the change of the branches can only occur at those singular points, if any). However, since this is just a transformation function, one does not care for making it as general as possible. On the contrary, its simplest form is most valuable as soon as the goal of the transformation, that is the wave equation simplification, is achieved. Our choice of $\chi(\Phi)$ can always provide it non-decreasing, i. e., with no extrema whatsoever. Its derivative turns to zero at most at the inflection points where $\omega(\Phi)=-3 /2$, if such points exist. We have thus transformed from a description in terms of the scalar field $\Phi$ and the two arbitrary functions $\omega(\Phi)$ and $\tilde{\Lambda}(\F)$, satisfying conditions (\ref{eq:om}) and (\ref{eq:lam}), respectively, to a description in terms of the scalar field $\chi$ and the two arbitrary functions $\Phi(\chi)$ and $\Lambda(\chi)$. The scalar field equation (\ref{eq:chi}) is a wave equation with a potential $ \Lambda(\chi)$. The source for the scalar field equation is proportional to the trace of the Einstein frame energy-momentum tensor. The factor of proportionality describes the coupling between matter and scalar field; whith our choice of transformation, this factor is positive for $0<\Phi<\infty$ (by (\ref{eq:defchi}), the derivative $d\ln\Phi/d\chi$ can only turn to zero when $\Phi \rightarrow\+0$ or $\Phi\rightarrow\infty$, and only if $\omega(\Phi)$ tends to infinity in the corresponding limit). Generally, the factor is scalar field dependent; the special case when it does not depend on the scalar field corresponds to $\omega=const$, i. e., to the Brans-Dicke theory. \section{Conclusions} We constructed the proper energy-momentum tensor of the scalar field in scalar-tensor gravity, disentangling simultaneously the dynamics of the scalar field from that of gravity {\it per se}. The Einstein frame arises naturally out of this disentanglement . We have shown that {\it all} scalar field terms on the right of the Einstein equation (\ref{eq:tG}) {\it cannot be identified} with the energy-momentum tensor of the scalar field because some of them contain the second covariant derivatives. The latter originate from variation of the gravitational part of the action, $\Phi \tilde{R}$, after an integration by parts. Hence they form a part of the dynamical description of gravity, and not of the scalar field. They occur because the dynamics of gravity and that of the purely scalar excitations are entangled in the physical frame, as a result of the nonminimal coupling between gravity and the scalar field. We defined a new connection in terms of which the full dynamics of the gravitational part can be explicitly separated. When doing this, one immediately finds the correct energy-momentum tensor of the scalar field as given by equation (\ref{eq:Sig}). This is a well defined energy-momentum tensor as long as we impose the condition that the energy density never becomes negative. This condition leads to the inequalities (\ref{eq:om}) and (\ref{eq:lam}) for the two otherwise arbitrary functions specifying the theory, the coupling function and the scalar field dependent cosmological "constant". We regard these restrictions as necessary for any ST theory to be physical. It is of importance that the dynamical connection can also be obtained by a conformal transformation of the metric. The conformally transformed metric is the geometrical object that describes the full dynamics of the tensor part of gravity. The Einstein conformal frame is the one defined in terms of the dynamical metric. Usually, the scalar field is redefined in the conformal frame, to simplify the scalar field equation in it. We analyze the behavior of the appropriate transformation function and demonstrate that it can always be chosen monotonic. Our hope is that the results presented here shed some light on the subject of the relationship between the physical and Einstein frames, which seems to have caused not a small amount of confusion before. In particular, the behavior of different quantities in each frame, such as the scalar field and its energy-momentum tensor, becomes clear, as well as the relation between the physical and dynamical metrics, and the correct dynamical description of gravity in scalar-tensor theories. \section*{Acknowledgments} This work was supported by NASA grant NAS 8-39225 to Gravity Probe~B. We are grateful to R.V.Wagoner for many valuable comments and to the Gravity Probe B Theory Group for fruitful discussions.